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 31 毫秒
1.
The most common mutation in cystic fibrosis (CF) is ΔF508, which is associated with failure of the mutant cystic fibrosis transmembrane conductance regulator (CFTR) to traffic to the plasma membrane. By a still unknown mechanism, the loss of correctly trafficked ΔF508-CFTR results in an excess of the epithelial sodium channel (ENaC) on the apical plasma membrane. ENaC trafficking is known to be regulated by a signaling pathway involving the glucocorticoid receptor, the serum- and glucocorticoid-regulated kinase SGK1, and the ubiquitin E3 ligase Nedd4-2. We show here that dexamethasone rescues functional expression of ΔF508-CFTR. The half-life of ΔF508-CFTR is also dramatically enhanced. Dexamethasone-activated ΔF508-CFTR rescue is blocked either by the glucocorticoid receptor antagonist RU38486 or by the phosphatidylinositol 3-kinase inhibitor LY294002. Co-immunoprecipitation studies indicate that Nedd4-2 binds to both wild-type- and ΔF508-CFTR. These complexes are inhibited by dexamethasone treatment, and CFTR ubiquitination is concomitantly decreased. We further show that knockdown of Nedd4-2 by small interfering RNA also corrects ΔF508-CFTR trafficking. Conversely, knockdown of SGK1 by small interfering RNA completely blocks dexamethasone-activated ΔF508-CFTR rescue. These data suggest that the SGK1/Nedd4-2 signaling pathway regulates both CFTR and ENaC trafficking in CF epithelial cells.Cystic fibrosis (CF)2 is the most common life-limiting genetic disease in the United States and is due to mutations in the CFTR gene. The most common mutation, ΔF508-CFTR, results in a failure of the mutant protein to traffic properly to the apical plasma membrane of epithelial cells in the lung and other organs (1, 2). By a still unknown mechanism, the loss of correctly trafficked ΔF508-CFTR results in an excess of the epithelial sodium channel (ENaC) on the apical plasma membrane (35). In the CF lung, such high levels of ENaC activity are believed to cause dehydration of the airway, and the consequent proinflammatory condition that characterizes CF lung pathophysiology. Similar proinflammatory pathophysiology has been reported to characterize the lung of transgenic mice which overexpress β-ENaC (6). Operationally, it seems that when membrane-localized CFTR decreases in CF, ENaC activity at the plasma membrane increases; CF-related morbidity and mortality follow.In the case of ENaC trafficking, the process is known to be regulated by a glucocorticoid receptor/SGK1 signaling pathway affecting phosphorylation of the ubiquitin ligase E3 protein Nedd4-2 (7, 8). Fig. 1 illustrates how surface expression of ENaC is controlled by the serum- and glucocorticoid-inducible kinase SGK1, the upstream signal, and the ubiquitin E3 ligase Nedd4-2, the downstream signal. Under default conditions, Nedd4-2 suppresses ENaC surface expression by binding to ENaC via the interaction between the PPXY motifs of ENaC and WW domains on Nedd4-2. Nedd4-2 then catalyzes the ubiquitination of bound ENaC. This step targets ENaC for proteasomal degradation (9, 10). However, when Nedd4-2 is phosphorylated by SGK1, the default interaction between Nedd4-2 and ENaC is reduced, and ENaC is maintained at the plasma membrane (7, 8). The requirement for Nedd4-2 for destruction of ENaC is supported by the recent observation that siRNA against Nedd4-2 is sufficient to permit ENaC to be expressed at the plasma membrane (10). Importantly, both glucocorticoid receptor (GR) and phosphoinositide-3-kinase (PI 3-kinase) signaling pathways must be present for high levels of Na+ transport to occur. For example, treatment with the GR antagonist RU38486 (1113) or the PI 3-kinase inhibitor LY294002 (1416) results in a complete loss of glucocorticoid-activated ENaC activity.Open in a separate windowFIGURE 1.Schematic diagram of regulation of ENaC and CFTR by SGK1/Nedd4-2. The surface expression of ENaC is controlled by the serum/glucocorticoid inducible kinase SGK1, the upstream signal, and the neural precursor cell-expressed developmentally down-regulated isoform 2 (Nedd4-2), the downstream signal. The solid black arrows trace the signal to a point where phospho-Nedd4-2 releases ENaC, thereby saving it from default ubiquitination and proteasomal destruction. ENaC is then maintained at the plasma membrane. Glucocorticoid-activated ENaC membrane trafficking is blocked by the glucocorticoid receptor antagonist RU38486 and the PI 3-kinase inhibitor LY294002. Alternatively, silencing of endogenous Nedd4-2 by siRNA enhances ENaC trafficking to the plasma membrane. (+) indicates positive regulation, and (−) indicates negative regulation.The placement of the parenthetical (CFTR) in the SGK1/Nedd4-2 signaling pathway (Fig. 1) serves to underscore our hypothesis that CFTR itself could play an interactive or parallel role in the SGK1/Nedd4-2/ENaC-trafficking mechanism. This hypothesis seems reasonable because the regulatory effects of SGK1 and Nedd4-2 are not limited to trafficking of ENaC but also regulate several other epithelial channels and transporters (17, 18). Additionally, co-expression studies in Xenopus oocytes (19, 20) have shown that SGK1 appears to greatly enhance the functional activity of CFTR.In this report we have shown that activation of the SGK1 signaling pathway by the glucocorticoid dexamethasone results in the rescue of ΔF508-CFTR. The half-life of ΔF508-CFTR, once it reaches the plasma membrane, is also dramatically enhanced. Consistently, glucocorticoid-activated ΔF508-CFTR rescue is blocked by the GR antagonist RU38486 and by the PI 3-kinase inhibitor LY294002 as well as by knockdown of endogenous SGK1 by siRNA. We have further shown that at the downstream end of the SGK1/Nedd4-2 signaling pathway, knockdown of Nedd4-2 by siRNA also results in ΔF508-CFTR rescue. Finally, co-immunoprecipitation studies indicated that Nedd4-2 binds to both WT- and ΔF508-CFTR and that treatment with either glucocorticoid or Nedd4-2 siRNA reduces formation of Nedd4-2·CFTR complexes as well as ubiquitination of ΔF508-CFTR. Consistently, chloride transport is well correlated with the level of plasma membrane expression of ΔF508-CFTR protein. These data suggest that the glucocorticoid receptor-dependent SGK1/Nedd4-2 signaling pathway regulates both CFTR and ENaC trafficking in CF epithelial cells. We interpret these results to indicate that drugs affecting the SGK1/Nedd4-2 signaling pathway may be promising targets for cystic fibrosis therapeutic development.  相似文献   

2.
Cystic Fibrosis (CF) is an autosomal recessive disorder caused by mutations in the gene encoding the Cystic fibrosis transmembrane conductance regulator (CFTR). ΔF508-CFTR, the most common disease-causing CF mutant, exhibits folding and trafficking defects and is retained in the endoplasmic reticulum, where it is targeted for proteasomal degradation. To identify signaling pathways involved in ΔF508-CFTR rescue, we screened a library of endoribonuclease-prepared short interfering RNAs (esiRNAs) that target ∼750 different kinases and associated signaling proteins. We identified 20 novel suppressors of ΔF508-CFTR maturation, including the FGFR1. These were subsequently validated by measuring channel activity by the YFP halide-sensitive assay following shRNA-mediated knockdown, immunoblotting for the mature (band C) ΔF508-CFTR and measuring the amount of surface ΔF508-CFTR by ELISA. The role of FGFR signaling on ΔF508-CFTR trafficking was further elucidated by knocking down FGFRs and their downstream signaling proteins: Erk1/2, Akt, PLCγ-1, and FRS2. Interestingly, inhibition of FGFR1 with SU5402 administered to intestinal organoids (mini-guts) generated from the ileum of ΔF508-CFTR homozygous mice resulted in a robust ΔF508-CFTR rescue. Moreover, combination of SU5402 and VX-809 treatments in cells led to an additive enhancement of ΔF508-CFTR rescue, suggesting these compounds operate by different mechanisms. Chaperone array analysis on human bronchial epithelial cells harvested from ΔF508/ΔF508-CFTR transplant patients treated with SU5402 identified altered expression of several chaperones, an effect validated by their overexpression or knockdown experiments. We propose that FGFR signaling regulates specific chaperones that control ΔF508-CFTR maturation, and suggest that FGFRs may serve as important targets for therapeutic intervention for the treatment of CF.Cystic fibrosis (CF)1 is a pleiotropic disease caused by an abnormal ion transport in the secretory epithelia lining the tubular organs of the body such as lungs, intestines, pancreas, liver, and male reproductive tract. In the airways of CF patients, reduced Cl and bicarbonate secretion caused by lack of functional Cystic fibrosis transmembrane conductance regulator (CFTR) on the apical surface, and hyper-absorption of Na+ because of elevated activity of ENaC (1), lead to a dehydration of the airway surface liquid (ASL). This reduces the viscosity of the mucus layer and the deposited layer of thickened mucus creates an environment that promotes bacterial colonization, which eventually leads to chronic infection of the lungs and death (2, 3).CFTR is a transmembrane protein that functions as a cAMP-regulated, ATP-dependent Cl channel that also allows passage of bicarbonate through its pore (4, 5). It also possesses ATPase activity important for Cl conductance (6, 7). The CFTR structure is predicted to consist of five domains: two membrane spanning domains (MSD1, MSD2), each composed of six putative transmembrane helices, two nucleotide binding domains (NBD1, NBD2), and a unique regulatory (R) region (8).More than 1900 CFTR mutations have been identified to date (www.genet.sickkids.on.ca/cftr). The most common mutation is a deletion of phenylalanine at position 508 (ΔF508 or ΔF508-CFTR) in NBD1 (9). The ΔF508 mutation causes severe defects in the processing and function of CFTR. The protein exhibits impaired trafficking from the endoplasmic reticulum (ER) to the plasma membrane (PM), impaired intramolecular interactions between NBD1 and the transmembrane domain, and cell surface instability (1015). Nevertheless, the ΔF508 defect can be corrected, because treating cells expressing ΔF508-CFTR with low temperature or chemical chaperones (e.g. glycerol) can restore some surface expression of the mutant (11, 16).Numerous small molecules that can at least partially correct (or potentiate) the ΔF508-CFTR defect have been identified to date (1727), and some were already tested in clinical trials (e.g. sildenafil, VX-809/Lumacaftor), or have made it to the clinic (VX-770/Kalydeco/Ivacaftor) (http://www.cff.org/research/DrugDevelopmentPipeline/). However, the need to identify new ΔF508-CFTR correctors remains immense as the most promising corrector, VX-809, has proven ineffective in alleviating lung disease of CF patients when administered alone (27). Thus, our group developed a high-content technology aimed at identifying proteins and small molecules that correct the trafficking and functional defects of ΔF508-CFTR (28). We successfully used this approach to carry out three separate high-content screens: a protein overexpression screen (28), a small-molecule kinase inhibitor screen (29) and a kinome RNA interference (RNAi) screen, described here.  相似文献   

3.
4.
Cannabis sativa is well known to produce unique secondary metabolites called cannabinoids. We recently discovered that Cannabis leaves induce cell death by secreting tetrahydrocannabinolic acid (THCA) into leaf tissues. Examinations using isolated Cannabis mitochondria demonstrated that THCA causes mitochondrial permeability transition (MPT) though opening of MPT pores, resulting in mitochondrial dysfunction (the important feature of necrosis). Although Ca2+ is known to cause opening of animal MPT pores, THCA directly opened Cannabis MPT pores in the absence of Ca2+. Based on these results, we conclude that THCA has the ability to induce necrosis though MPT in Cannabis leaves, independently of Ca2+. We confirmed that other cannabinoids (cannabidiolic acid and cannabigerolic acid) also have MPT-inducing activity similar to that of THCA. Moreover, mitochondria of plants which do not produce cannabinoids were shown to induce MPT by THCA treatment, thus suggesting that many higher plants may have systems to cause THCA-dependent necrosis.Key words: cannabinoid, Cannabis sativa, cylophilin D, mitochondrial permeability transition, necrosisCannabis sativa produces unique secondary metabolites consisting of alkylresorcinol and monoterpene groups.1 These metabolites called cannabinoids are well known to show a variety of interesting pharmacological activities including psychoactive effect and analgesic effect. Therefore, cannabinoids have attracted a great deal of attention, whereas why C. sativa produces such metabolites has long remained unclear. However, we have recently obtained evidences indicating the physiological function of THCA in Cannabis leaves.2We discovered that THCA is stored in capitate-sessile glands on Cannabis leaves and that secretion of this cannabinoid into leaf tissues causes cell death. When the properties of THCA were examined using cultured Cannabis cells, this cannabinoid induced plasmamembrane shrinkage and DNA degradation. These responses are regarded as the features of apoptotic cells, but were not suppressed by apoptosis inhibitors. In contrast, the necrosis inhibitor cyclosporine A significantly inhibited both plasmamembrane shrinkage and DNA degradation in Cannabis cells. Therefore, we assumed that THCA induces necrotic cell death in Cannabis cells and leaves.Necrosis in plants and animals is usually triggered by MPT though opening of MPT pores.3,4 MPT is known to cause mitochondrial dysfunction by mitochondrial swelling and loss of mitochondrial membrane potential (ΔΨm),5,6 and we also confirmed that THCA induces mitochondrial swelling and ΔΨm reduction in mitochondria isolated from Cannabis cells and that pretreatment with cyclosporine A inhibits both responses. Based on these evidences, we concluded that THCA has the activity to induce MPT-dependent necrosis.As described above, MPT pores play an important role in necrosis induction, whereas the mechanism of their opening in higher plants has not been fully understood. However, binding of cyclophilin D (a protein present in mitochondrial matrix) to MPT pores is shown to be essential for their opening in plants as well as animal.79 In animal mitochondria, Ca2+ mediates this binding reaction, leading to opening of MPT pores. Wheat mitochondria are also shown to undergo swelling through opening of MPT pores in response to Ca2+,9 whereas MPT pores of oats,10 Arabidopsis thaliana11 and C. sativa2 do not open by Ca2+ treatment. In contrast, THCA catalyzed opening of Cannabis MPT pores in the absence of Ca2+, suggesting that THCA directly mediates binding of cyclophilin D to MPT pores (Fig. 1). In addition, we have now confirmed that THCA causes dysfunction though MPT in mitochondria of plants (rice, soybean, A. thaliana and Scutellaria baicalensis) lacking cannabinoid-producing ability (data not shown). Therefore, many higher plants may have the systems to induce THCA-dependent necrosis.Open in a separate windowFigure 1A model depicting the opening mechanism of MPT pores in mitochondria. CYD, cyclophilin D; CN, cannabinoid.Furthermore, we investigated whether other cannabinoids and their related compounds can mediate MPT in Cannabis mitochondria. When the MPT-inducing activity of each sample was measured by monitoring both ΔΨm reduction (Fig. 2) and mitochondrial swelling (data not shown), we confirmed that cannabinoids tested here (cannabidiolic acid and cannabigerolic acid) possess the activities similar to those of THCA. On the other hand, olivetolic acid (the akylresorcinol moiety of cannabinoid) and geraniol (the monoterpene moiety of cannabigerolic acid) showed neither ΔΨm reduction nor mitochondrial swelling (Fig. 2). These results suggested that the structures (cannabinoid skeleton) where monoterpene and olivetolic acid are coupled to each other seem essential for opening of MPT pores. Therefore, we assumed that plant cyclophilin D and MPT pores have the cannabinoid-binding site.Open in a separate windowFigure 2Change of ΔΨm by treatment with various compounds (A) and their chemical structures (B). The isolated mitochondria were stained with the ΔΨm-indicating reagent (tetramethylrhodamine methylester, TMRM) and then incubated with 200 µM of each compound for 60 min. The intensity of TMRM fluorescence was measured using a fluorescence microplate reader. A decrease of the fluorescence intensity indicates ΔΨm reduction. CBDA, cannabidiolic acid; CBGA, cannabigerolic acid; OLA, olivetolic acid.Plant cell death is shown to participate in important physiological responses such as leaf senescence, somatic embryogenesis and defense against microbial pathogens.12,13 Based on its induction mechanism, plant cell death is largely classified into apoptosis and necrosis. Although the molecular mechanism of apoptosis has been extensively investigated, there is little precise information on plant necrosis. However, our study would provide important insight into necrosis-inducing mechanisms in higher plants.  相似文献   

5.
Cell surface receptors of the integrin family are pivotal to cell adhesion and migration. The activation state of heterodimeric αβ integrins is correlated to the association state of the single-pass α and β transmembrane domains. The association of integrin αIIbβ3 transmembrane domains, resulting in an inactive receptor, is characterized by the asymmetric arrangement of a straight (αIIb) and tilted (β3) helix relative to the membrane in congruence to the dissociated structures. This allows for a continuous association interface centered on helix-helix glycine-packing and an unusual αIIb(GFF) structural motif that packs the conserved Phe-Phe residues against the β3 transmembrane helix, enabling αIIb(D723)β3(R995) electrostatic interactions. The transmembrane complex is further stabilized by the inactive ectodomain, thereby coupling its association state to the ectodomain conformation. In combination with recently determined structures of an inactive integrin ectodomain and an activating talin/β complex that overlap with the αβ transmembrane complex, a comprehensive picture of integrin bi-directional transmembrane signaling has emerged.Key words: cell adhesion, membrane protein, integrin, platelet, transmembrane complex, transmembrane signalingThe communication of biological signals across the plasma membrane is fundamental to cellular function. The ubiquitous family of integrin adhesion receptors exhibits the unusual ability to convey signals bi-directionally (outside-in and inside-out signaling), thereby controlling cell adhesion, migration and differentiation.15 Integrins are Type I heterodimeric receptors that consist of large extracellular domains (>700 residues), single-pass transmembrane (TM) domains, and mostly short cytosolic tails (<70 residues). The activation state of heterodimeric integrins is correlated to the association state of the TM domains of their α and β subunits.610 TM dissociation initiated from the outside results in the transmittal of a signal into the cell, whereas dissociation originating on the inside results in activation of the integrin to bind ligands such as extracellular matrix proteins. The elucidation of the role of the TM domains in integrin-mediated adhesion and signaling has been the subject of extensive research efforts, perhaps commencing with the demonstration that the highly conserved GFFKR sequence motif of α subunits (Fig. 1), which closely follows the first charged residue on the intracellular face, αIIb(K989), constrains the receptor to a default low affinity state.11 Despite these efforts, an understanding of this sequence motif had not been reached until such time as the structure of the αIIb TM segment was determined.12 In combination with the structure of the β3 TM segment13 and available mutagenesis data,6,9,10,14,15 this has allowed the first correct prediction of the overall association of an integrin αβ TM complex.12 The predicted association was subsequently confirmed by the αIIbβ3 complex structure determined in phospholipid bicelles,16 as well as by the report of a similar structure based on molecular modeling using disulfide-based structural constraints.17 In addition to the structures of the dissociated and associated αβ TM domains, their membrane embedding was defined12,13,16,18,19 and it was experimentally recognized that, in the context of the native receptor, the TM complex is stabilized by the inactive, resting ectodomain.16 These advances in integrin membrane structural biology are complemented by the recent structures of a resting integrin ectodomain and an activating talin/β cytosolic tail complex that overlap with the αβ TM complex,20,21 allowing detailed insight into integrin bi-directional TM signaling.Open in a separate windowFigure 1Amino acid sequence of integrin αIIb and β3 transmembrane segments and flanking regions. Membrane-embedded residues12,13,16,18,19 are enclosed by a gray box. Residues 991–995 constitute the highly conserved GFFKR sequence motif of integrin α subunits.  相似文献   

6.
7.
8.
9.
10.
11.
Characterization of aggregation profiles of monoclonal antibodies (mAb) is gaining importance because an increasing number of mAb-based therapeutics are entering clinical studies and gaining marketing approval. To develop a successful formulation, it is imperative to identify the critical biochemical properties of each potential mAb drug candidate. We investigated the conformational change and aggregation of a human IgG1 using external dye-binding experiments with fluorescence spectroscopy and compared the aggregation profiles obtained to the results of size-exclusion chromatography. We show that using an appropriate dye at selected mAb concentration, unfolding or aggregation can be studied. In addition, dye-binding experiments may be used as conventional assays to study therapeutic mAb stability.Key words: therapeutic monoclonal antibody, protein aggregation, conformational change, stability and shelf-life prediction, accelerated studiesMonoclonal antibodies (mAbs) have emerged as a novel class of protein drugs and are utilized for a variety of mostly incurable and debilitating diseases such as cancer and rheumatoid arthritis.14 For treatment of chronic diseases, it is desirable for these drugs to be administered subcutaneously, in which case high protein concentrations (>100 mg/mL) are generally needed.5,6 Protein-based drugs containing mAbs must contain minimum amounts of aggregation and fragmentation and conserve their structural integrity during storage because degraded or aggregated protein may induce immunogenicity or reduce efficacy. Currently, size-exclusion chromatography-high performance liquid chromatography (SEC-HPLC) is the most commonly used method to characterize mAb aggregation profiles;7 however it is time consuming, expensive and requires expertise. SEC-HPLC cannot be used to obtain accurate biophysical profiles of mAbs at high concentrations because dilution during the experiment might lead to reversible aggregation. Furthermore, the potential interaction of aggregates with surfaces, e.g., needle, tubing, column, will lead to the loss of sample and thus an inaccurate analysis.8,9 Additional drawbacks of the technique are that different conformations such as partially unfolded monomers also cannot be distinguished by SEC-HPLC and large aggregates may be totally excluded during the injection into the column.External dye binding assays have been used to characterize protein stability and aggregation,1012 and studies involving biopharmaceuticals have been reported recently, e.g., for thermostability screening10 and detection of aggregation.1114 These methods are not limited by protein quantity and are more sensitive because they are fluorescence-based. We studied the accelerated unfolding of an IgG1 mAb with the hydrophobic dye 1-anilino-8-naphthale-nesulfonate (ANS), and its accelerated aggregation with aggregate specific Thioflavin T (ThT). We have also conducted accelerated aggregation studies with SEC-HPLC7 and compared the findings to the ThT binding results. We hypothesize that key structures formed during mAb aggregation can be probed selectively by the appropriate dyes (Fig. 1) with specific mAb concentrations.Open in a separate windowFigure 1Key structures of the mAb probed by fluorescent dyes. N and U are native and unfolded monomers, respectively. “n” reactive monomers form aggregates.  相似文献   

12.
The dynamic remodeling of actin filaments in guard cells functions in stomatal movement regulation. In our previous study, we found that the stochastic dynamics of guard cell actin filaments play a role in chloroplast movement during stomatal movement. In our present study, we further found that tubular actin filaments were present in tobacco guard cells that express GFP-mouse talin; approximately 2.3 tubular structures per cell with a diameter and height in the range of 1–3 µm and 3–5 µm, respectively. Most of the tubular structures were found to be localized in the cytoplasm near the inner walls of the guard cells. Moreover, the tubular actin filaments altered their localization slowly in the guard cells of static stoma, but showed obvious remodeling, such as breakdown and re-formation, in moving guard cells. Tubular actin filaments were further found to be colocalized with the chloroplasts in guard cells, but their roles in stomatal movement regulation requires further investigation.Key words: actin dynamics, tubular actin filaments, chloroplast, guard cell, stomatal movementStomatal movement responses to surrounding environment are mediated by guard cell signaling.1,2 Actin filaments within guard cells are dynamic cytoarchitectures and function in stomatal development and movement.3 Arrays of actin filaments in guard cells that are dependent on different stomatal apertures have also been reported in references 47. For example, the random or longitudinal orientations of actin filaments in closed stomata change to a radial orientation or ring-like array after stomata opening.5,6,8 The reorganization of the actin architecture during stomatal movement depends on the depolymerization and repolymerization of actin filaments in guard cells. In contrast to the traditional treadmill model of actin dynamic mechanisms, stochastic dynamics of actin have been revealed in plant cells, such as in the epidermal cells of hypocotyl and root, the pavement cells of Arabidopsis cotyledons, and the guard cells of tobacco (Nicotiana tabacum).911 In this alternative system, the short actin fragments generated from severed long filaments can link with each other to form longer filaments by end-joining activity. The actin regulatory proteins, Arp2/3 complex, capping protein and actin depolymerizing factor (ADF)/cofilin, may also be involved in the stochastic dynamics of actin filaments.12,13Using tobacco GFP-mouse talin expression lines, we have previously analyzed the stochastic dynamics of guard cell actin filaments and their roles in chloroplast displacement during stomatal movement.6,11 We found from these analyses that another arrangement of actin filaments, i.e., tubular actin filaments, exists in the guard cells of these tobacco lines. We first found the circle-like actin filaments in 82% of the guard cells (counting 320 cells) in tobacco expressing GFPmouse talin when analyzing a single optical section (Fig. 1A). In a previous study of BY-2 cells expressing GFP-Lifeact labeled actin filaments, Smertenko et al. found similar structures, i.e., quoit-like structures or acquosomes in all of the plant tissues examined except growing root hairs.10 However, in our present analysis of serial sections, we determined that the circle-like actin filaments in the tobacco guard cells were long tubes (Fig. 1A), as the lengths (about 3–5 µm) of these structures were greater than their diameter (about 1–3 µm). Hence, we denoted these structures as tubular actin filaments to distinguish them from the circular conformations of actin filaments observed previously in other plant cell tissues.10,1419 About 2.3 of these tubular actin filaments were found per guard cell, which is less than the number of acquosomes reported in BY-2 cells (about 6.7 per cell).10 Analysis of serial optical sections at the z-axis revealed that the tubular actin filaments localize in the cytoplasm near the inner walls of the guard cells (Fig. 1B), which is similar to the distribution of chloroplasts in guard cells.11 Longitudinal sections further revealed a colocalization of tubular actin filaments and chloroplasts (Fig. 1B).Open in a separate windowFigure 1Tubular actin filaments in the guard cells of a tobacco (Nicotiana tabacum) line expressing GFP-mouse talin. (A) Optical-sections (interval, 1.5 µm) of guard cells in a moving stoma showing tubular actin filaments (arrow heads). Frames (a1) and (a2) are cross sections of 1.5-µm-picture through the yellow and red lines, respectively, revealing the cross section of the circle structures are parallel lines (arrows). (B) Optical-sections of a stoma from the outer periclinal walls to the inner walls of the guard cells (interval, 1 µm). The tubular actin filaments (arrow heads) are localized in the cytoplasm near to the inner periclinal walls of guard cells. Frame (b1) is the guard cell on the right of the frame “4 µm”; (b2) is the cross section of b1 through the red line; and (b3) is a higher magnification image of the area encompassed by the white square in b2. Arrows indicate the colocalization between the tubular actin filaments and the chloroplast (indicated using a red pseudocolor). (C) Time-series imaging showing the movement of tubular actin filaments in the guard cells of static stomata. Frame (c1) comprises three images colored red (0 S), green (40 S) and blue (80 S), that are merged in a single frame to show the translocation of the tubular actin filaments (arrows). (D) Time-series images of the opening stomata showing the breakdown (arrows) and re-formation (arrowheads) of the tubular actin filaments. All images were captured using a Zeiss LSM 510 META confocal laser scanning microscope, as described by Wang et al.11 Bars, 10 µm.We performed time-lapse imaging and found that the translocation of tubular actin filaments is slow in static stomata in which the distance between two tubular actin filaments typically increased from 2.22 to 2.50 µm after 80 sec (Fig. 1C). In moving stomata, however, the tubular actin filaments showed an obvious dynamic reorganization whereby they could be processed into short fragments and also reemerged after they had disintegrated (Fig. 1D). These results indicate that tubular actin filaments have stochastic dynamics that are similar to the long actin filaments of guard cells.11 In our previous study, we found that the stochastic dynamics of actin filaments correlate with light-induced chloroplast movement in guard cells.11 However, whether the dynamics of the tubular actin filaments are also involved in chloroplast movement during stomatal movement remains to be investigated. In cultured mesophyll cells which had been mechanically isolated from Zinnia elegans, Wilsen et al. previously found a close association between fully closed actin rings and chloroplasts.18 These authors further found that the average percentage of cells with free actin rings increased at the initial culture stage, and then decreased, which indicates that the formation of actin rings might be a response of the actin cytoskeleton to cellular stress or disturbance.18 The turgor pressure of guard cells is the fundamental basis of stomatal movement leading to changes in the shape, volume, wall structure, and membrane surface of guard cells.2024 We speculate from our current data that there is a relationship between tubular actin filaments and the shape changes of guard cells during stomatal movement.  相似文献   

13.
Environmental and developmental signals can elicit differential activation of membrane proton (H+) fluxes as one of the primary responses of plant and fungal cells. In recent work,1 we could determine that during the presymbiotic growth of arbuscular mycorrhizal (AM) fungi specific domains of H+ flux are activated by clover root factors, namely host root exudates or whole root system. Consequently, activation on hyphal growth and branching were observed and the role of plasma membrane H+-ATPase was investigated. The specific inhibitors differentially abolished most of hyphal H+ effluxes and fungal growth. As this enzyme can act in signal transduction pathways, we believe that spatial and temporal oscillations of the hyphal H+ fluxes could represent a pH signature for both early events of the AM symbiosis and fungal ontogeny.Key words: H+-specific vibrating probe, pH signatures, arbuscular mycorrhiza, pH signalling, Gigaspora margaritaThe 450-million-year-old symbiosis between the majority of land plants and arbuscular mycorrhizal (AM) fungi is one of the most ancient, abundant and ecologically important symbiosis on Earth.2,3The development of AM interaction starts before the physical contact between the host plant roots and the AM fungus. The hyphal growth and branching are induced by the root factors exudated by host plants, followed by the formation of appressorium leading to the hyphal penetration in the root system. These root factors seems to be specifically synthesized by host plants, since exudates from non-host plants are not able to promote neither hyphal differentiation nor appressorium formation.4,5 Most root exudates contain several host signals or better, active compounds including flavonoids6,19 and strigolactones,7,8 however many of them are not yet known.Protons (H+) may have an important role on the fungal growth and host signal perception.1 In plant and fungal cells, H+ can be pumped out through two different mechanisms: (1) the activity of the P-type plasma membrane (PM) H+-ATPase9 and (2) PM redox reactions.10 The proportional contribution from both mechanisms is not known, but in most plant cells the PM H+-ATPase seems to be the major responsible by the H+ efflux across plasma membrane. AM Fungal cells also energize their PM using P-type H+-pumps quite similar to the plant ones. Indeed, some genes codifying isoforms of P-type H+-ATPase have been isolated of AM fungi,1113 and AM fungal ATP hydrolysis activity was shown by cytochemistry, localized mainly in the first 70 µm from the germ tube tip.14 This structural evidence correlates with data obtained by H+-specific vibrating probe (Fig. 1A and B), which indicates that the H+ efflux in Gigaspora margarita is more intense in the subapical region of the lateral hyphae1 (Fig. 1A). Furthermore, the correlation between the cytosolic pH profile previously obtained by Jolicoeur et al.,15 with the H+ efflux pattern (erythrosine-dependent), seems to clearly indicate that an active PM H+-ATPase takes place at the subapical hyphal region. Using orthovanadate, we could show that those H+ effluxes are susceptible mainly in the subapical region, but no effect in the apical was found.1 Recently, a method to use fluorescent marker expression in an AM fungus driven by arbuscular mycorrhizal promoters was published.31 It could be adjusted as an alternative to measure “in vivo” PM H+-ATPase expression in AM fungal hyphae and their responses to root factors.31Open in a separate windowFigure 1(A) H+ flux profile along growing secondary hyphae of G. margarita in the presence (open squares) or absence (closed squares) of erythrosin B and its correlation with cytosolic pH (pHc) data described by Jolicoeur et al.,15 (dotted line). Dotted area depicts the region with higher susceptibility to erythrosin B. (B) ion-selective electrode near to AM fungal hyphae. (C) Stimulation on hyphal H+ efflux after incubation with root factors or whole root system. R, roots; RE, root exudates; CO2, carbon dioxide; CWP, cell wall proteins; GR24, synthetic strigolactone. The medium pH in all treatment was monitored and remained about 5.7, including with prior CO2 incubation. Means followed by the same letter are statistically equal by Duncan''s test at p < 5%.The H+ electrochemical gradient generated by PM H+-ATPases provides not only driving force for nutrient uptake,9,16 but also can act as an intermediate in signal transduction pathways.18 The participation of these H+ pumps in cell polarity and tip growth of plant cells was recently reported,27 addressing their crucial role on apical growth.28 Naturally, in the absence of root factors the AM fungi have basal metabolic8,2123 and respiratory activity.24 However when root signals are recognized and processed by AM fungal cells they might become activated.22 We thus searched for pH signatures that could reflect the alterations on fungal metabolism in response to external stimuli. In fact, preliminary analyses from our group demonstrate that AM fungal hyphae increase their H+ efflux in response not only to root exudates recognition, but also to other root factors (Fig. 1C). The incubation for 30 min of AM fungal hyphae with several root factors induces hyphal H+ efflux similar to the response to intact root system (5 days of incubation). The major increases were found with 1% CO2 (750%) followed by root cell wall proteins (221%), root exudates (130%) and synthetic strigolactone (5%) (Fig. 1C). Those stimulations could define the transition from the state without root signals to the presymbiotic developmental stage (Fig. 1C). In the case of CO2, the incorporation of additional carbon could represent a new source of energy, since CO2 dark fixation takes place in Glomus intraradices germ tubes.22,25Interestingly, after the treatment with synthetic strigolactone (10−5 M GR24), no significant stimulation was found compared to the remaining factors (Fig. 1C). It opens the question if the real effect of strigolactone is restrict to hyphal branching and does not intervene in very fast response pathways. Likewise, strigolactones need additional time to exhibit an effect, as recently discussed by Steinkellner et al.,26 However, at the moment, no comprehensive electrophysiological analyses are presently available separating the effects of strigolactone and some flavonoids in AM fungal hyphae.The next target of our work is the study of ionic responses of single germ tubes or primary hyphae to root factors (Fig. 2). As reported by Ramos et al.,1 we have been observing that the pattern of ion fluxes at the apical zone of primary hyphae is differentiated from secondary or lateral hyphae. In the primary, two interesting responses were detected in the absence of root factors: (1) a “dormant Ca2+ flux” and (2) Cl or anion fluxes at the same direction of H+ ions, suggesting a possible presence of H+/Cl symporters at the apex, similarly to what occurs in root hairs (Fig. 2).30 In the presence of root factors such as root exudates the stimulated influxes of Cl (anion), H+, Na+ and effluxes of K+ and Ca2+ are activated. It can explain why the AM fungi hyphal tips are depolarized20,29 during the period without root signals—“asymbiosis”—as long as K+ efflux and H+ influx occur simultaneously. Indeed, H+ as well as Ca2+ ions may act as second messengers, where extra and intracellular transient pH changes are preconditions for a number of processes, including gravity responses and possibly in plant-microbe interactions.17,30Open in a separate windowFigure 2Ion dynamics in the apex of primary hyphae of arbuscular mycorrhizal fungi. It represents the Stage 1 described in Ramos et al.1 After treatment with root factors, an activation of Ca2+ efflux is observed at the hyphal apex.Clearly, further data on the mechanism of action of signaling molecules such as strigolactones over the signal transduction and ion dynamics in AM fungi will be very important to improve our understanding of the molecular bases of the mycorrhization process. Future studies are necessary in order to provide basic knowledge of the ion signaling mechanisms and their role on the response of very important molecules playing at the early events of AM symbiosis.  相似文献   

14.
Organelle movement in plants is dependent on actin filaments with most of the organelles being transported along the actin cables by class XI myosins. Although chloroplast movement is also actin filament-dependent, a potential role of myosin motors in this process is poorly understood. Interestingly, chloroplasts can move in any direction and change the direction within short time periods, suggesting that chloroplasts use the newly formed actin filaments rather than preexisting actin cables. Furthermore, the data on myosin gene knockouts and knockdowns in Arabidopsis and tobacco do not support myosins'' XI role in chloroplast movement. Our recent studies revealed that chloroplast movement and positioning are mediated by the short actin filaments localized at chloroplast periphery (cp-actin filaments) rather than cytoplasmic actin cables. The accumulation of cp-actin filaments depends on kinesin-like proteins, KAC1 and KAC2, as well as on a chloroplast outer membrane protein CHUP1. We propose that plants evolved a myosin XI-independent mechanism of the actin-based chloroplast movement that is distinct from the mechanism used by other organelles.Key words: actin, Arabidopsis, blue light, kinesin, myosin, organelle movement, phototropinOrganelle movement and positioning are pivotal aspects of the intracellular dynamics in most eukaryotes. Although plants are sessile organisms, their organelles are quickly repositioned in response to fluctuating environmental conditions and certain endogenous signals. By and large, plant organelle movements and positioning are dependent on actin filaments, although microtubules play certain accessory roles in organelle dynamics.1,2 Actin inhibitors effectively retard the movements of mitochondria,36 peroxisomes,5,711 Golgi stacks,12,13 endoplasmic reticulum (ER),14,15 and nuclei.1618 These organelles are co-aligned and associated with actin filaments.5,7,8,1012,15,18 Recent progress in this field started to reveal the molecular motility system responsible for the organelle transport in plants.19Chloroplast movement is among the most fascinating models of organelle movement in plants because it is precisely controlled by ambient light conditions.20,21 Weak light induces chloroplast accumulation response so that chloroplasts can capture photosynthetic light efficiently (Fig. 1A). Strong light induces chloroplast avoidance response to escape from photodamage (Fig. 1B).22 The blue light-induced chloroplast movement is mediated by the blue light receptor phototropin (phot). In some cryptogam plants, the red light-induced chloroplast movement is regulated by a chimeric phytochrome/phototropin photoreceptor neochrome.2325 In a model plant Arabidopsis, phot1 and phot2 function redundantly to regulate the accumulation response,26 whereas phot2 alone is essential for the avoidance response.27,28 Several additional factors regulating chloroplast movement were identified by analyses of Arabidopsis mutants deficient in chloroplast photorelocation.2932 In particular, identification of CHUP1 (chloroplast unusual positioning 1) revealed the connection between chloroplasts and actin filaments at the molecular level.29 CHUP1 is a chloroplast outer membrane protein capable of interacting with F-actin, G-actin and profilin in vitro.29,33,34 The chup1 mutant plants are defective in both the chloroplast movement and chloroplast anchorage to the plasma membrane,22,29,33 suggesting that CHUP1 plays an important role in linking chloroplasts to the plasma membrane through the actin filaments. However, how chloroplasts move using the actin filaments and whether chloroplast movement utilizes the actin-based motility system similar to other organelle movements remained to be determined.Open in a separate windowFigure 1Schematic distribution patterns of chloroplasts in a palisade cell under different light conditions, weak (A) and strong (B) lights. Shown as a side view of mid-part of the cell and a top view with three different levels (i.e., top, middle and bottom of the cell). The cell was irradiated from the leaf surface shown as arrows. Weak light induces chloroplast accumulation response (A) and strong light induces the avoidance response (B).Here, we review the recent findings pointing to existence of a novel actin-based mechanisms for chloroplast movement and discuss the differences between the mechanism responsible for movement of chloroplasts and other organelles.  相似文献   

15.
16.
Microglia engage in the clearance of dead cells or dangerous debris. When neighboring cells are injured, the cells release or leak ATP into extracellular space and microglia rapidly move toward or extend a process to the nucleotides as chemotaxis through P2Y12 receptors. In the meanwhile, microglia express the metabotropic P2Y6 receptors, the activation of which by uridine 5′-diphosphate (UDP) triggers microglial phagocytosis in a concentration-dependent fashion. UDP/UTP was leaked when hippocampal neurons were damaged by kainic acid in vivo and in vitro. Systemic administration of kainic acid in rats resulted in neuronal cell death in the hippocampal CA1 and CA3 regions, where increases in mRNA for P2Y6 receptors in activated microglia. Thus, the P2Y6 receptor is upregulated when neurons are damaged, and would function as a sensor for phagocytosis by sensing diffusible UDP signals.Key Words: microglia, phagocytosis, P2Y6 receptors, UDPAccumulating findings indicate that nucleotides play an important role in neuron to glia communication through P2 purinoceptors, even though ATP is recognized primarily to be a source of free energy and nucleotides are key molecules in cells. P2 purinoceptors are divided into two families, ionotropic receptors (P2X) and metabotropic receptors (P2Y) (Fig. 1). P2X receptors (seven types; P2X1-P2X7) contain intrinsic pores that open by binding with ATP. P2Y (eight types; P2Y1,2,4,6 and 11–14) are activated by nucleotides and couple to intracellular second-messenger systems through heteromeric G-proteins.1 Microglia express P2X4, P2X7, P2Y2, P2Y6 and P2Y121 and are known as resident macrophages in CNS, accounting for 5–10% of the total population of glia.2,3 When neurons are injured or dead, microglia are activated, resulting in their interaction with immune cells, active migration to the site of injury, release of pro-inflammatory substances and the phagocytosis of damaged cells or debris. For such activation of microglial motilities, extracellular nucleotides have a central role. Extracellular ATP functions as a chemoattractant. Microglial chemotaxis by ATP via P2Y12 receptors was originally found by Honda et al.,4 and has recently been confirmed in vivo in P2Y12 receptor knockout animals.5 Neuronal injury results in the release or leakage of ATP that appears to be a “find-me” signal from damaged neurons to microglia to cause chemotaxis. In addition to microglial migration by ATP, another nucleotide, UDP, an endogenous agonist of the P2Y6 receptor, greatly activates the motility of microglia and orders microglia to engulf damaged neurons.6Open in a separate windowFigure 1P2 purinergic receptors (ATP receptors).Phagocytosis is a specialized form of endocytosis taking relatively large particles (> 1.0 µm) into vacuoles and has a central role in tissue remodeling, inflammation and the defense against infectious agents.7 Phagocytosis is initiated by the activation of cell-surface phagocytosis receptors, including Fc receptors, complement receptors, integrins, endotoxin receptors (CD18, CD14), mannose receptors and scavenger receptors8 which are activated by corresponding extracellular ligands called as “eat-me” signals. Since recognition is the most important step for phagocytosis, extensive studies on phagocytosis receptors have been reported. With regard to apoptotic cells, it is well known that dying cells express so called “eat-me” signals such as phosphatidylserine (PS) on their surface membrane,8 by which microglia recognize the apoptotic cells in order to catch and remove them.8 As for amyloid β protein (Aβ), a key molecule that mediates Alzheimer''s disease, microglia remove Aβ presumably via Fc receptor-dependent phagocytosis.9,10 It, however, is unclear how phagocytotic cells come to the target cells or debris. Our findings suggest that nucleotides might be the molecules to guide phagocytotic cells to the targets.We found that exogenously applied UDP caused microglial phagocytosis through P2Y6 in a concentration-dependent manner, and that neuronal injury caused by kainic acid (KA) upregulated P2Y6 receptors in microglia, the KA evoked neuronal injury resulted in an increase in extracellular UTP, which was immediately metabolized into UDP in vivo and in vitro. We also found that UDP leaked from injured neurons caused P2Y6 receptor-dependent phagocytosis in vivo and in vitro. Thus, UDP could be a diffusible molecule that signals the crisis of damaged neurons to microglia, triggering phagocytosis. Nucleotides seem to have the ability to act as “eat-us” signals for necrotic cells suffering traumatic or ischemic injury because such necrotic cells cause swelling, followed by shrinkage, leading to the leakage of cytoplasmic molecules including a large amount of ATP and UTP and extracellular nucleotides are immediately degraded by ecto-nucleotideases, suggesting that leaked nucleotides could be transient and localized signals that alert to the crisis created by the presence of the necrotic cells. These findings suggest that microglia might be attracted by ATP/ADP4,5,11,12 and subsequently recognize UDP, starting to recognize “eat-me” signals attached to the targets and engulf them (Fig. 2). It is interesting that ATP/ADP is not able to efficiently activate P2Y6 receptors, nor can UDP act on P2Y12 receptors. Thus, adenine and uridine nucleotides would regulate microglial motilities, i.e. chemotaxis and phagocytosis, in a coordinated fashion.Open in a separate windowFigure 2Illustration of nucleotide-activated microglial chemotaxix and phagocytosis. Activated microglia might be attracted by ATP/ADP is not able to efficiently activate P2Y6 receptors, nor ca UDP act on P2Y12 receptors.  相似文献   

17.
18.
Processes putatively dependent on the galactolipid monogalactosyldiacylglycerol (MGDG) were recently studied using the knockdown monogalactosyldiacylglycerol synthase 1 (mgd1-1) mutant (∼40% reduction in MGDG). Surprisingly, targeting of chloroplast proteins was not affected in mgd1-1 mutants, suggesting they retain sufficient MGDG to maintain efficient targeting. However, in dark-grown mgd1-1 plants the photoactive to photoinactive protochlorophyllide (Pchlide) ratio was increased, suggesting that photoprotective responses are induced in them. Nevertheless, mgd1-1 could not withstand high light intensities, apparently due to impairment of another photoprotective mechanism, the xanthophyll cycle (and hence thermal dissipation). This was mediated by increased conductivity of the thylakoid membrane leading to a higher pH in the thylakoid interior, which impaired the pH-dependent activation of violaxanthin de-epoxidase (VDE) and PsbS. These findings suggest that MGDG contribute directly to the regulation of photosynthesis-related processes.Key words: conductivity, galactolipid, light stress, photosynthesis, plastid, xanthophyllThe galactolipid monogalactosyldiacylglycerol (MGDG), the major lipid in plastids,1 is mainly synthesised in inner plastid envelopes,2 where monogalactosyldiacylglycerol synthase 1 (MGD1) catalyses the last step of its production.3 Two MGDG-deficient mutants are known: the knockdown mgd1-1 mutant, which accumulates ∼40% less MGDG than wild type,4 and the null mutant mgd1-2, which displays extremely severe defects in chloroplast and plant development.5 Thus, the mgd1-1 mutant is more suitable for assessing putative roles of MGDG in processes such as protein targeting and photoprotection.There are conflicting indications regarding the involvement of galactolipids in chloroplast protein targeting: some suggest they play an important role,610 but not all.11,12 The data recently collected for mgd1-1 do not support MGDG''s involvement in protein targeting, since (inter alia) the level of MGDG in mgd1-1 mutants is clearly sufficient for efficient targeting.13 Further, the galactolipid associated with the TOC complex12 is digalactosyldiacylglycerol (DGDG) and the digalactosyldiacylglycerol synthase 1 (dgd1) mutant,14 which has ∼10% of wild-type levels of DGDG, has impaired import efficiency.15,16 Hence, this may indicate that DGDG is relatively more important for chloroplast import than MGDG.The prolamellar bodies (PLBs) of etioplasts have high lipid-to-protein ratios compared to thylakoids. Their major lipid and protein are MGDG and NADPH:Pchlide oxidoreductase (POR), respectively,17 and MGDG putatively plays an important role, interactively with POR, in the formation of PLBs.1820 The transformation of PLBs into thylakoids involves phototransformation of photoactive Pchlide (F656), a precursor of chlorophyll. Non-photoactive Pchlide (F631) is susceptible to photooxidative damage, but POR is believed to suppress this.21,22 After excitation at 440 nm, mgd1-1 mutants display distinctly higher fluorescence emission peaks corresponding to photoactive Pchlide than wild type counterparts and (hence) higher photoactive:non-photoactive Pchlide ratios.13 These changes may be photoprotective responses that favour formation of photoactive Pchlide and optimize the plants'' opportunities to use light for chlorophyll production, enabling the transformation of etioplasts into chloroplasts.Interestingly,the xanthophyll cycle, another photoprotective mechanism, is impaired in mgd1-1.13 Normally, the xanthophyll cycle pigment violaxanthin is de-epoxidized into antheraxanthin, and then into zeaxanthin, by the enzyme VDE (Fig. 1), which is dependent on MGDG.23 MGDG is also an integral component of photosynthetic complexes.2426 Thus, since mgd1-1 mutants have reduced total amounts of xanthophyll and chlorophyll pigments, but increased chlorophyll a/b ratios, their photosynthesis capacity is unsurprisingly reduced, even though the organization of their electron transport chains is not strongly affected by the MGDG deficiency.13Open in a separate windowFigure 1Reactions of the xanthophyll cycle (adapted from ref. 29). VDE, violaxanthin de-epoxidase; ZE, zeaxanthin epoxidase.During short-term high light stress, antheraxanthin and zeaxanthin are thought to facilitate dissipation of excess light energy in the PSII antenna bed by non-photochemical quenching.27,28 Upon high light stress the pH decreases, triggering photoprotective mechanisms via changes in the PSII antenna system. The PsbS protein, which is involved in thermal dissipation, is protonated and initiates a conformational change in the PSII antenna bed. This change is further stabilized by the de-epoxidation of violaxanthin to zeaxanthin by the luminal VDE.28 However, the thermal dissipation is impaired in mgd1-1 mutants at high light intensities (>1000 µmol m−2 s−1) making them more susceptible to light stress. Surprisingly, this is not mediated by direct effects on VDE and PsbS activities, but by changes in the proton conductivity of the thylakoid membrane.13The steady-state capacity of the xanthophyll cycle is reduced in mgd1-1 mutants, due to a ∼40% reduction in the proton motive force (pmf) across their thylakoid membranes, indicating that they have impaired capacities to energize these membranes. Nevertheless, the pmf is more or less equal to wild type under light-limited conditions (200 µmol m−2 s−1 light); it is only the increase in pmf in high light intensities that is impaired in the mutants.13 This leads to the thylakoid lumen being less acidic in mgd1-1 than in wild type, hampering full activation of VDE and PsbS. Thus, the thylakoid lumen pH is above the threshold level required for full activation of PsbS and VDE under steady-state conditions and so de-epoxidation rates are retarded and the equilibrium between zeaxanthin and violaxanthin starts to shift slightly towards violaxanthin (Fig. 2).13 Thus, increased conductivity of the thylakoid membranes is probably responsible for the diminished non-photochemical quenching in mgd1-1, and the findings strongly indicate that MGDG is required for efficient photosynthesis and photoprotection, in addition to being a physical membrane constituent.Open in a separate windowFigure 2Schematic diagram illustrating the normal mode of action of the xanthophyll cycle. In standard light conditions, V is bound to the photosynthetic complexes and harvests light. In strong light, V is released from the complexes and converted to Z by VDE, which is unable to access V when it is associated with the photosynthetic complexes. The newly formed Z then binds to the photosynthetic complexes (at the PsbS protein), where it dissipates excess energy through NPQ. V, violaxanthin; A, antheraxanthin; Z, zeaxanthin; VDE, violaxanthin de-epoxidase; ZE, zeaxanthin epoxidase. Arrows indicate the directions of reactions.  相似文献   

19.
20.
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