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.
NHE5 is a brain-enriched Na+/H+ exchanger that dynamically shuttles between the plasma membrane and recycling endosomes, serving as a mechanism that acutely controls the local pH environment. In the current study we show that secretory carrier membrane proteins (SCAMPs), a group of tetraspanning integral membrane proteins that reside in multiple secretory and endocytic organelles, bind to NHE5 and co-localize predominantly in the recycling endosomes. In vitro protein-protein interaction assays revealed that NHE5 directly binds to the N- and C-terminal cytosolic extensions of SCAMP2. Heterologous expression of SCAMP2 but not SCAMP5 increased cell-surface abundance as well as transporter activity of NHE5 across the plasma membrane. Expression of a deletion mutant lacking the SCAMP2-specific N-terminal cytosolic domain, and a mini-gene encoding the N-terminal extension, reduced the transporter activity. Although both Arf6 and Rab11 positively regulate NHE5 cell-surface targeting and NHE5 activity across the plasma membrane, SCAMP2-mediated surface targeting of NHE5 was reversed by dominant-negative Arf6 but not by dominant-negative Rab11. Together, these results suggest that SCAMP2 regulates NHE5 transit through recycling endosomes and promotes its surface targeting in an Arf6-dependent manner.Neurons and glial cells in the central and peripheral nervous systems are especially sensitive to perturbations of pH (1). Many voltage- and ligand-gated ion channels that control membrane excitability are sensitive to changes in cellular pH (1-3). Neurotransmitter release and uptake are also influenced by cellular and organellar pH (4, 5). Moreover, the intra- and extracellular pH of both neurons and glia are modulated in a highly transient and localized manner by neuronal activity (6, 7). Thus, neurons and glia require sophisticated mechanisms to finely tune ion and pH homeostasis to maintain their normal functions.Na+/H+ exchangers (NHEs)3 were originally identified as a class of plasma membrane-bound ion transporters that exchange extracellular Na+ for intracellular H+, and thereby regulate cellular pH and volume. Since the discovery of NHE1 as the first mammalian NHE (8), eight additional isoforms (NHE2-9) that share 25-70% amino acid identity have been isolated in mammals (9, 10). NHE1-5 commonly exhibit transporter activity across the plasma membrane, whereas NHE6-9 are mostly found in organelle membranes and are believed to regulate organellar pH in most cell types at steady state (11). More recently, NHE10 was identified in human and mouse osteoclasts (12, 13). However, the cDNA encoding NHE10 shares only a low degree of sequence similarity with other known members of the NHE gene family, raising the possibility that this sodium-proton exchanger may belong to a separate gene family distantly related to NHE1-9 (see Ref. 9).NHE gene family members contain 12 putative transmembrane domains at the N terminus followed by a C-terminal cytosolic extension that plays a role in regulation of the transporter activity by protein-protein interactions and phosphorylation. NHEs have been shown to regulate the pH environment of synaptic nerve terminals and to regulate the release of neurotransmitters from multiple neuronal populations (14-16). The importance of NHEs in brain function is further exemplified by the findings that spontaneous or directed mutations of the ubiquitously expressed NHE1 gene lead to the progression of epileptic seizures, ataxia, and increased mortality in mice (17, 18). The progression of the disease phenotype is associated with loss of specific neuron populations and increased neuronal excitability. However, NHE1-null mice appear to develop normally until 2 weeks after birth when symptoms begin to appear. Therefore, other mechanisms may compensate for the loss of NHE1 during early development and play a protective role in the surviving neurons after the onset of the disease phenotype.NHE5 was identified as a unique member of the NHE gene family whose mRNA is expressed almost exclusively in the brain (19, 20), although more recent studies have suggested that NHE5 might be functional in other cell types such as sperm (21, 22) and osteosarcoma cells (23). Curiously, mutations found in several forms of congenital neurological disorders such as spinocerebellar ataxia type 4 (24-26) and autosomal dominant cerebellar ataxia (27-29) have been mapped to chromosome 16q22.1, a region containing NHE5. However, much remains unknown as to the molecular regulation of NHE5 and its role in brain function.Very few if any proteins work in isolation. Therefore identification and characterization of binding proteins often reveal novel functions and regulation mechanisms of the protein of interest. To begin to elucidate the biological role of NHE5, we have started to explore NHE5-binding proteins. Previously, β-arrestins, multifunctional scaffold proteins that play a key role in desensitization of G-protein-coupled receptors, were shown to directly bind to NHE5 and promote its endocytosis (30). This study demonstrated that NHE5 trafficking between endosomes and the plasma membrane is regulated by protein-protein interactions with scaffold proteins. More recently, we demonstrated that receptor for activated C-kinase 1 (RACK1), a scaffold protein that links signaling molecules such as activated protein kinase C, integrins, and Src kinase (31), directly interacts with and activates NHE5 via integrin-dependent and independent pathways (32). These results further indicate that NHE5 is partly associated with focal adhesions and that its targeting to the specialized microdomain of the plasma membrane may be regulated by various signaling pathways.Secretory carrier membrane proteins (SCAMPs) are a family of evolutionarily conserved tetra-spanning integral membrane proteins. SCAMPs are found in multiple organelles such as the Golgi apparatus, trans-Golgi network, recycling endosomes, synaptic vesicles, and the plasma membrane (33, 34) and have been shown to play a role in exocytosis (35-38) and endocytosis (39). Currently, five isoforms of SCAMP have been identified in mammals. The extended N terminus of SCAMP1-3 contain multiple Asn-Pro-Phe (NPF) repeats, which may allow these isoforms to participate in clathrin coat assembly and vesicle budding by binding to Eps15 homology (EH)-domain proteins (40, 41). Further, SCAMP2 was shown recently to bind to the small GTPase Arf6 (38), which is believed to participate in traffic between the recycling endosomes and the cell surface (42, 43). More recent studies have suggested that SCAMPs bind to organellar membrane type NHE7 (44) and the serotonin transporter SERT (45) and facilitate targeting of these integral membrane proteins to specific intracellular compartments. We show in the current study that SCAMP2 binds to NHE5, facilitates the cell-surface targeting of NHE5, and elevates Na+/H+ exchange activity at the plasma membrane, whereas expression of a SCAMP2 deletion mutant lacking the N-terminal domain containing the NPF repeats suppresses the effect. Further we show that this activity of SCAMP2 requires an active form of a small GTPase Arf6, but not Rab11. We propose a model in which SCAMPs bind to NHE5 in the endosomal compartment and control its cell-surface abundance via an Arf6-dependent pathway.  相似文献   

2.
3.
4.
5.
The Ca2+ release-activated Ca2+ channel is a principal regulator of intracellular Ca2+ rise, which conducts various biological functions, including immune responses. This channel, involved in store-operated Ca2+ influx, is believed to be composed of at least two major components. Orai1 has a putative channel pore and locates in the plasma membrane, and STIM1 is a sensor for luminal Ca2+ store depletion in the endoplasmic reticulum membrane. Here we have purified the FLAG-fused Orai1 protein, determined its tetrameric stoichiometry, and reconstructed its three-dimensional structure at 21-Å resolution from 3681 automatically selected particle images, taken with an electron microscope. This first structural depiction of a member of the Orai family shows an elongated teardrop-shape 150Å in height and 95Å in width. Antibody decoration and volume estimation from the amino acid sequence indicate that the widest transmembrane domain is located between the round extracellular domain and the tapered cytoplasmic domain. The cytoplasmic length of 100Å is sufficient for direct association with STIM1. Orifices close to the extracellular and intracellular membrane surfaces of Orai1 seem to connect outside the molecule to large internal cavities.Ca2+ is an intracellular second messenger that plays important roles in various physiological functions such as immune response, muscle contraction, neurotransmitter release, and cell proliferation. Intracellular Ca2+ is mainly stored in the endoplasmic reticulum (ER).2 This ER system is distributed through the cytoplasm from around the nucleus to the cell periphery close to the plasma membrane. In non-excitable cells, the ER releases Ca2+ through the inositol 1,4,5-trisphosphate (IP3) receptor channel in response to various signals, and the Ca2+ store is depleted. Depletion of Ca2+ then induces Ca2+ influx from outside the cell to help in refilling the Ca2+ stores and to continue Ca2+ rise for several minutes in the cytoplasm (1, 2). This Ca2+ influx was first proposed by Putney (3) and was named store-operated Ca2+ influx. In the immune system, store-operated Ca2+ influx is mainly mediated by the Ca2+ release-activated Ca2+ (CRAC) current, which is a highly Ca2+-selective inwardly rectified current with low conductance (4, 5). Pathologically, the loss of CRAC current in T cells causes severe combined immunodeficiency (6) where many Ca2+ signal-dependent gene expressions, including cytokines, are interrupted (7). Therefore, CRAC current is necessary for T cell functions.Recently, Orai1 (also called CRACM1) and STIM1 have been physiologically characterized as essential components of the CRAC channel (812). They are separately located in the plasma membrane and in the ER membrane; co-expression of these proteins presents heterologous CRAC-like currents in various types of cells (10, 1315). Both of them are shown to be expressed ubiquitously in various tissues (1618). STIM1 senses Ca2+ depletion in the ER through its EF hand motif (19) and transmits a signal to Orai1 in the plasma membrane. Although Orai1 is proposed as a regulatory component for some transient receptor potential canonical channels (20, 21), it is believed from the mutation analyses to be the pore-forming subunit of the CRAC channel (8, 2224). In the steady state, both Orai1 and STIM1 molecules are dispersed in each membrane. When store depletion occurs, STIM1 proteins gather into clusters to form puncta in the ER membrane near the plasma membrane (11, 19). These clusters then trigger the clustering of Orai1 in the plasma membrane sites opposite the puncta (25, 26), and CRAC channels are activated (27).Orai1 has two homologous genes, Orai2 and Orai3 (8). They form the Orai family and have in common the four transmembrane (TM) segments with relatively large N and C termini. These termini are demonstrated to be in the cytoplasm, because both N- and C-terminally introduced tags are immunologically detected only in the membrane-permeabilized cells (8, 9). The subunit stoichiometry of Orai1 is as yet controversial: it is believed to be an oligomer, presumably a dimer or tetramer even in the steady state (16, 2830).Despite the accumulation of biochemical and electrophysiological data, structural information about Orai1 is limited due to difficulties in purification and crystallization. In this study, we have purified Orai1 in its tetrameric form and have reconstructed the three-dimensional structure from negatively stained electron microscopic (EM) images.  相似文献   

6.
In Alzheimer disease (AD), frontotemporal dementia and parkinsonism linked to chromosome 17 (FTDP-17) and other tauopathies, tau accumulates and forms paired helical filaments (PHFs) in the brain. Tau isolated from PHFs is phosphorylated at a number of sites, migrates as ∼60-, 64-, and 68-kDa bands on SDS-gel, and does not promote microtubule assembly. Upon dephosphorylation, the PHF-tau migrates as ∼50–60-kDa bands on SDS-gels in a manner similar to tau that is isolated from normal brain and promotes microtubule assembly. The site(s) that inhibits microtubule assembly-promoting activity when phosphorylated in the diseased brain is not known. In this study, when tau was phosphorylated by Cdk5 in vitro, its mobility shifted from ∼60-kDa bands to ∼64- and 68-kDa bands in a time-dependent manner. This mobility shift correlated with phosphorylation at Ser202, and Ser202 phosphorylation inhibited tau microtubule-assembly promoting activity. When several tau point mutants were analyzed, G272V, P301L, V337M, and R406W mutations associated with FTDP-17, but not nonspecific mutations S214A and S262A, promoted Ser202 phosphorylation and mobility shift to a ∼68-kDa band. Furthermore, Ser202 phosphorylation inhibited the microtubule assembly-promoting activity of FTDP-17 mutants more than of WT. Our data indicate that FTDP-17 missense mutations, by promoting phosphorylation at Ser202, inhibit the microtubule assembly-promoting activity of tau in vitro, suggesting that Ser202 phosphorylation plays a major role in the development of NFT pathology in AD and related tauopathies.Neurofibrillary tangles (NFTs)4 and senile plaques are the two characteristic neuropathological lesions found in the brains of patients suffering from Alzheimer disease (AD). The major fibrous component of NFTs are paired helical filaments (PHFs) (for reviews see Refs. 13). Initially, PHFs were found to be composed of a protein component referred to as “A68” (4). Biochemical analysis reveled that A68 is identical to the microtubule-associated protein, tau (4, 5). Some characteristic features of tau isolated from PHFs (PHF-tau) are that it is abnormally hyperphosphorylated (phosphorylated on more sites than the normal brain tau), does not bind to microtubules, and does not promote microtubule assembly in vitro. Upon dephosphorylation, PHF-tau regains its ability to bind to and promote microtubule assembly (6, 7). Tau hyperphosphorylation is suggested to cause microtubule instability and PHF formation, leading to NFT pathology in the brain (13).PHF-tau is phosphorylated on at least 21 proline-directed and non-proline-directed sites (8, 9). The individual contribution of these sites in converting tau to PHFs is not entirely clear. However, some sites are only partially phosphorylated in PHFs (8), whereas phosphorylation on specific sites inhibits the microtubule assembly-promoting activity of tau (6, 10). These observations suggest that phosphorylation on a few sites may be responsible and sufficient for causing tau dysfunction in AD.Tau purified from the human brain migrates as ∼50–60-kDa bands on SDS-gel due to the presence of six isoforms that are phosphorylated to different extents (2). PHF-tau isolated from AD brain, on the other hand, displays ∼60-, 64-, and 68 kDa-bands on an SDS-gel (4, 5, 11). Studies have shown that ∼64- and 68-kDa tau bands (the authors have described the ∼68-kDa tau band as an ∼69-kDa band in these studies) are present only in brain areas affected by NFT degeneration (12, 13). Their amount is correlated with the NFT densities at the affected brain regions. Moreover, the increase in the amount of ∼64- and 68-kDa band tau in the brain correlated with a decline in the intellectual status of the patient. The ∼64- and 68-kDa tau bands were suggested to be the pathological marker of AD (12, 13). Biochemical analyses determined that ∼64- and 68-kDa bands are hyperphosphorylated tau, which upon dephosphorylation, migrated as normal tau on SDS-gel (4, 5, 11). Tau sites involved in the tau mobility shift to ∼64- and 68-kDa bands were suggested to have a role in AD pathology (12, 13). It is not known whether phosphorylation at all 21 PHF-sites is required for the tau mobility shift in AD. However, in vitro the tau mobility shift on SDS-gel is sensitive to phosphorylation only on some sites (6, 14). It is therefore possible that in the AD brain, phosphorylation on some sites also causes a tau mobility shift. Identification of such sites will significantly enhance our knowledge of how NFT pathology develops in the brain.PHFs are also the major component of NFTs found in the brains of patients suffering from a group of neurodegenerative disorders collectively called tauopathies (2, 11). These disorders include frontotemporal dementia and Parkinsonism linked to chromosome 17 (FTDP-17), corticobasal degeneration, progressive supranuclear palsy, and Pick disease. Each PHF-tau isolated from autopsied brains of patients suffering from various tauopathies is hyperphosphorylated, displays ∼60-, 64-, and 68-kDa bands on SDS-gel, and is incapable of binding to microtubules. Upon dephosphorylation, the above referenced PHF-tau migrates as a normal tau on SDS-gel, binds to microtubules, and promotes microtubule assembly (2, 11). These observations suggest that the mechanisms of NFT pathology in various tauopathies may be similar and the phosphorylation-dependent mobility shift of tau on SDS-gel may be an indicator of the disease. The tau gene is mutated in familial FTDP-17, and these mutations accelerate NFT pathology in the brain (1518). Understanding how FTDP-17 mutations promote tau phosphorylation can provide a better understanding of how NFT pathology develops in AD and various tauopathies. However, when expressed in CHO cells, G272V, R406W, V337M, and P301L tau mutations reduce tau phosphorylation (19, 20). In COS cells, although G272V, P301L, and V337M mutations do not show any significant affect, the R406W mutation caused a reduction in tau phosphorylation (21, 22). When expressed in SH-SY5Y cells subsequently differentiated into neurons, the R406W, P301L, and V337M mutations reduce tau phosphorylation (23). In contrast, in hippocampal neurons, R406W increases tau phosphorylation (24). When phosphorylated by recombinant GSK3β in vitro, the P301L and V337M mutations do not have any effect, and the R406W mutation inhibits phosphorylation (25). However, when incubated with rat brain extract, all of the G272V, P301L, V337M, and R406W mutations stimulate tau phosphorylation (26). The mechanism by which FTDP-17 mutations promote tau phosphorylation leading to development of NFT pathology has remained unclear.Cyclin-dependent protein kinase 5 (Cdk5) is one of the major kinases that phosphorylates tau in the brain (27, 28). In this study, to determine how FTDP-17 missense mutations affect tau phosphorylation, we phosphorylated four FTDP-17 tau mutants (G272V, P301L, V337M, and R406W) by Cdk5. We have found that phosphorylation of tau by Cdk5 causes a tau mobility shift to ∼64- and 68 kDa-bands. Although the mobility shift to a ∼64-kDa band is achieved by phosphorylation at Ser396/404 or Ser202, the mobility shift to a 68-kDa band occurs only in response to phosphorylation at Ser202. We show that in vitro, FTDP-17 missense mutations, by promoting phosphorylation at Ser202, enhance the mobility shift to ∼64- and 68-kDa bands and inhibit the microtubule assembly-promoting activity of tau. Our data suggest that Ser202 phosphorylation is the major event leading to NFT pathology in AD and related tauopathies.  相似文献   

7.
Hepatocellular carcinoma (HCC) is one of the most common and aggressive human malignancies. Recombinant tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) is a promising anti-tumor agent. However, many HCC cells show resistance to TRAIL-induced apoptosis. In this study, we showed that bortezomib, a proteasome inhibitor, overcame TRAIL resistance in HCC cells, including Huh-7, Hep3B, and Sk-Hep1. The combination of bortezomib and TRAIL restored the sensitivity of HCC cells to TRAIL-induced apoptosis. Comparing the molecular change in HCC cells treated with these agents, we found that down-regulation of phospho-Akt (P-Akt) played a key role in mediating TRAIL sensitization of bortezomib. The first evidence was that bortezomib down-regulated P-Akt in a dose- and time-dependent manner in TRAIL-treated HCC cells. Second, LY294002, a PI3K inhibitor, also sensitized resistant HCC cells to TRAIL-induced apoptosis. Third, knocking down Akt1 by small interference RNA also enhanced TRAIL-induced apoptosis in Huh-7 cells. Finally, ectopic expression of mutant Akt (constitutive active) in HCC cells abolished TRAIL sensitization effect of bortezomib. Moreover, okadaic acid, a protein phosphatase 2A (PP2A) inhibitor, reversed down-regulation of P-Akt in bortezomib-treated cells, and PP2A knockdown by small interference RNA also reduced apoptosis induced by the combination of TRAIL and bortezomib, indicating that PP2A may be important in mediating the effect of bortezomib on TRAIL sensitization. Together, bortezomib overcame TRAIL resistance at clinically achievable concentrations in hepatocellular carcinoma cells, and this effect is mediated at least partly via inhibition of the PI3K/Akt pathway.Hepatocellular carcinoma (HCC)2 is currently the fifth most common solid tumor worldwide and the fourth leading cause of cancer-related death. To date, surgery is still the only curative treatment but is only feasible in a small portion of patients (1). Drug treatment is the major therapy for patients with advanced stage disease. Unfortunately, the response rate to traditional chemotherapy for HCC patients is unsatisfactory (1). Novel pharmacological therapy is urgently needed for patients with advanced HCC. In this regard, the approval of sorafenib might open a new era of molecularly targeted therapy in the treatment of HCC patients.Tumor necrosis factor (TNF)-related apoptosis-inducing ligand (TRAIL), a type II transmembrane protein and a member of the TNF family, is a promising anti-tumor agent under clinical investigation (2). TRAIL functions by engaging its receptors expressed on the surface of target cells. Five receptors specific for TRAIL have been identified, including DR4/TRAIL-R1, DR5/TRAIL-R2, DcR1, DcR2, and osteoprotegerin. Among TRAIL receptors, only DR4 and DR5 contain an effective death domain that is essential to formation of death-inducing signaling complex (DISC), a critical step for TRAIL-induced apoptosis. Notably, the trimerization of the death domains recruits an adaptor molecule, Fas-associated protein with death domain (FADD), which subsequently recruits and activates caspase-8. In type I cells, activation of caspase-8 is sufficient to activate caspase-3 to induce apoptosis; however, in another type of cells (type II), the intrinsic mitochondrial pathway is essential for apoptosis characterized by cleavage of Bid and release of cytochrome c from mitochondria, which subsequently activates caspase-9 and caspase-3 (3).Although TRAIL induces apoptosis in malignant cells but sparing normal cells, some tumor cells are resistant to TRAIL-induced apoptosis. Mechanisms responsible for the resistance include receptors and intracellular resistance. Although the cell surface expression of DR4 or DR5 is absolutely required for TRAIL-induced apoptosis, tumor cells expressing these death receptors are not always sensitive to TRAIL due to intracellular mechanisms. For example, the cellular FLICE-inhibitory protein (c-FLIP), a homologue to caspase-8 but without protease activity, has been linked to TRAIL resistance in several studies (4, 5). In addition, inactivation of Bax, a proapoptotic Bcl-2 family protein, resulted in resistance to TRAIL in MMR-deficient tumors (6, 7), and reintroduction of Bax into Bax-deficient cells restored TRAIL sensitivity (8), indicating that the Bcl-2 family plays a critical role in intracellular mechanisms for resistance of TRAIL.Bortezomib, a proteasome inhibitor approved clinically for multiple myeloma and mantle cell lymphoma, has been investigated intensively for many types of cancer (9). Accumulating studies indicate that the combination of bortezomib and TRAIL overcomes the resistance to TRAIL in various types of cancer, including acute myeloid leukemia (4), lymphoma (1013), prostate (1417), colon (15, 18, 19), bladder (14, 16), renal cell carcinoma (20), thyroid (21), ovary (22), non-small cell lung (23, 24), sarcoma (25), and HCC (26, 27). Molecular targets responsible for the sensitizing effect of bortezomib on TRAIL-induced cell death include DR4 (14, 27), DR5 (14, 20, 2223, 28), c-FLIP (4, 11, 2123, 29), NF-κB (12, 24, 30), p21 (16, 21, 25), and p27 (25). In addition, Bcl-2 family also plays a role in the combinational effect of bortezomib and TRAIL, including Bcl-2 (10, 21), Bax (13, 22), Bak (27), Bcl-xL (21), Bik (18), and Bim (15).Recently, we have reported that Akt signaling is a major molecular determinant in bortezomib-induced apoptosis in HCC cells (31). In this study, we demonstrated that bortezomib overcame TRAIL resistance in HCC cells through inhibition of the PI3K/Akt pathway.  相似文献   

8.
Members of the CLC gene family either function as chloride channels or as anion/proton exchangers. The plant AtClC-a uses the pH gradient across the vacuolar membrane to accumulate the nutrient in this organelle. When AtClC-a was expressed in Xenopus oocytes, it mediated exchange and less efficiently mediated Cl/H+ exchange. Mutating the “gating glutamate” Glu-203 to alanine resulted in an uncoupled anion conductance that was larger for Cl than . Replacing the “proton glutamate” Glu-270 by alanine abolished currents. These could be restored by the uncoupling E203A mutation. Whereas mammalian endosomal ClC-4 and ClC-5 mediate stoichiometrically coupled 2Cl/H+ exchange, their transport is largely uncoupled from protons. By contrast, the AtClC-a-mediated accumulation in plant vacuoles requires tight coupling. Comparison of AtClC-a and ClC-5 sequences identified a proline in AtClC-a that is replaced by serine in all mammalian CLC isoforms. When this proline was mutated to serine (P160S), Cl/H+ exchange of AtClC-a proceeded as efficiently as exchange, suggesting a role of this residue in exchange. Indeed, when the corresponding serine of ClC-5 was replaced by proline, this Cl/H+ exchanger gained efficient coupling. When inserted into the model Torpedo chloride channel ClC-0, the equivalent mutation increased nitrate relative to chloride conductance. Hence, proline in the CLC pore signature sequence is important for exchange and conductance both in plants and mammals. Gating and proton glutamates play similar roles in bacterial, plant, and mammalian CLC anion/proton exchangers.CLC proteins are found in all phyla from bacteria to humans and either mediate electrogenic anion/proton exchange or function as chloride channels (1). In mammals, the roles of plasma membrane CLC Cl channels include transepithelial transport (25) and control of muscle excitability (6), whereas vesicular CLC exchangers may facilitate endocytosis (7) and lysosomal function (810) by electrically shunting vesicular proton pump currents (11). In the plant Arabidopsis thaliana, there are seven CLC isoforms (AtClC-a–AtClC-g)2 (1215), which may mostly reside in intracellular membranes. AtClC-a uses the pH gradient across the vacuolar membrane to transport the nutrient nitrate into that organelle (16). This secondary active transport requires a tightly coupled exchange. Astonishingly, however, mammalian ClC-4 and -5 and bacterial EcClC-1 (one of the two CLC isoforms in Escherichia coli) display tightly coupled Cl/H+ exchange, but anion flux is largely uncoupled from H+ when is transported (1721). The lack of appropriate expression systems for plant CLC transporters (12) has so far impeded structure-function analysis that may shed light on the ability of AtClC-a to perform efficient exchange. This dearth of data contrasts with the extensive mutagenesis work performed with CLC proteins from animals and bacteria.The crystal structure of bacterial CLC homologues (22, 23) and the investigation of mutants (17, 1921, 2429) have yielded important insights into their structure and function. CLC proteins form dimers with two largely independent permeation pathways (22, 25, 30, 31). Each of the monomers displays two anion binding sites (22). A third binding site is observed when a certain key glutamate residue, which is located halfway in the permeation pathway of almost all CLC proteins, is mutated to alanine (23). Mutating this gating glutamate in CLC Cl channels strongly affects or even completely suppresses single pore gating (23), whereas CLC exchangers are transformed by such mutations into pure anion conductances that are not coupled to proton transport (17, 19, 20). Another key glutamate, located at the cytoplasmic surface of the CLC monomer, seems to be a hallmark of CLC anion/proton exchangers. Mutating this proton glutamate to nontitratable amino acids uncouples anion transport from protons in the bacterial EcClC-1 protein (27) but seems to abolish transport altogether in mammalian ClC-4 and -5 (21). In those latter proteins, anion transport could be restored by additionally introducing an uncoupling mutation at the gating glutamate (21).The functional complementation by AtClC-c and -d (12, 32) of growth phenotypes of a yeast strain deleted for the single yeast CLC Gef1 (33) suggested that these plant CLC proteins function in anion transport but could not reveal details of their biophysical properties. We report here the first functional expression of a plant CLC in animal cells. Expression of wild-type (WT) and mutant AtClC-a in Xenopus oocytes indicate a general role of gating and proton glutamate residues in anion/proton coupling across different isoforms and species. We identified a proline in the CLC signature sequence of AtClC-a that plays a crucial role in exchange. Mutating it to serine, the residue present in mammalian CLC proteins at this position, rendered AtClC-a Cl/H+ exchange as efficient as exchange. Conversely, changing the corresponding serine of ClC-5 to proline converted it into an efficient exchanger. When proline replaced the critical serine in Torpedo ClC-0, the relative conductance of this model Cl channel was drastically increased, and “fast” protopore gating was slowed.  相似文献   

9.
10.
Mitochondrial dysregulation is strongly implicated in Parkinson disease. Mutations in PTEN-induced kinase 1 (PINK1) are associated with familial parkinsonism and neuropsychiatric disorders. Although overexpressed PINK1 is neuroprotective, less is known about neuronal responses to loss of PINK1 function. We found that stable knockdown of PINK1 induced mitochondrial fragmentation and autophagy in SH-SY5Y cells, which was reversed by the reintroduction of an RNA interference (RNAi)-resistant plasmid for PINK1. Moreover, stable or transient overexpression of wild-type PINK1 increased mitochondrial interconnectivity and suppressed toxin-induced autophagy/mitophagy. Mitochondrial oxidant production played an essential role in triggering mitochondrial fragmentation and autophagy in PINK1 shRNA lines. Autophagy/mitophagy served a protective role in limiting cell death, and overexpressing Parkin further enhanced this protective mitophagic response. The dominant negative Drp1 mutant inhibited both fission and mitophagy in PINK1-deficient cells. Interestingly, RNAi knockdown of autophagy proteins Atg7 and LC3/Atg8 also decreased mitochondrial fragmentation without affecting oxidative stress, suggesting active involvement of autophagy in morphologic remodeling of mitochondria for clearance. To summarize, loss of PINK1 function elicits oxidative stress and mitochondrial turnover coordinated by the autophagic and fission/fusion machineries. Furthermore, PINK1 and Parkin may cooperate through different mechanisms to maintain mitochondrial homeostasis.Parkinson disease is an age-related neurodegenerative disease that affects ∼1% of the population worldwide. The causes of sporadic cases are unknown, although mitochondrial or oxidative toxins such as 1-methyl-4-phenylpyridinium, 6-hydroxydopamine (6-OHDA),3 and rotenone reproduce features of the disease in animal and cell culture models (1). Abnormalities in mitochondrial respiration and increased oxidative stress are observed in cells and tissues from parkinsonian patients (2, 3), which also exhibit increased mitochondrial autophagy (4). Furthermore, mutations in parkinsonian genes affect oxidative stress response pathways and mitochondrial homeostasis (5). Thus, disruption of mitochondrial homeostasis represents a major factor implicated in the pathogenesis of sporadic and inherited parkinsonian disorders (PD).The PARK6 locus involved in autosomal recessive and early-onset PD encodes for PTEN-induced kinase 1 (PINK1) (6, 7). PINK1 is a cytosolic and mitochondrially localized 581-amino acid serine/threonine kinase that possesses an N-terminal mitochondrial targeting sequence (6, 8). The primary sequence also includes a putative transmembrane domain important for orientation of the PINK1 domain (8), a conserved kinase domain homologous to calcium calmodulin kinases, and a C-terminal domain that regulates autophosphorylation activity (9, 10). Overexpression of wild-type PINK1, but not its PD-associated mutants, protects against several toxic insults in neuronal cells (6, 11, 12). Mitochondrial targeting is necessary for some (13) but not all of the neuroprotective effects of PINK1 (14), implicating involvement of cytoplasmic targets that modulate mitochondrial pathobiology (8). PINK1 catalytic activity is necessary for its neuroprotective role, because a kinase-deficient K219M substitution in the ATP binding pocket of PINK1 abrogates its ability to protect neurons (14). Although PINK1 mutations do not seem to impair mitochondrial targeting, PD-associated mutations differentially destabilize the protein, resulting in loss of neuroprotective activities (13, 15).Recent studies indicate that PINK1 and Parkin interact genetically (3, 16-18) to prevent oxidative stress (19, 20) and regulate mitochondrial morphology (21). Primary cells derived from PINK1 mutant patients exhibit mitochondrial fragmentation with disorganized cristae, recapitulated by RNA interference studies in HeLa cells (3).Mitochondria are degraded by macroautophagy, a process involving sequestration of cytoplasmic cargo into membranous autophagic vacuoles (AVs) for delivery to lysosomes (22, 23). Interestingly, mitochondrial fission accompanies autophagic neurodegeneration elicited by the PD neurotoxin 6-OHDA (24, 25). Moreover, mitochondrial fragmentation and increased autophagy are observed in neurodegenerative diseases including Alzheimer and Parkinson diseases (4, 26-28). Although inclusion of mitochondria in autophagosomes was once believed to be a random process, as observed during starvation, studies involving hypoxia, mitochondrial damage, apoptotic stimuli, or limiting amounts of aerobic substrates in facultative anaerobes support the concept of selective mitochondrial autophagy (mitophagy) (29, 30). In particular, mitochondrially localized kinases may play an important role in models involving oxidative mitochondrial injury (25, 31, 32).Autophagy is involved in the clearance of protein aggregates (33-35) and normal regulation of axonal-synaptic morphology (36). Chronic disruption of lysosomal function results in accumulation of subtly impaired mitochondria with decreased calcium buffering capacity (37), implicating an important role for autophagy in mitochondrial homeostasis (37, 38). Recently, Parkin, which complements the effects of PINK1 deficiency on mitochondrial morphology (3), was found to promote autophagy of depolarized mitochondria (39). Conversely, Beclin 1-independent autophagy/mitophagy contributes to cell death elicited by the PD toxins 1-methyl-4-phenylpyridinium and 6-OHDA (25, 28, 31, 32), causing neurite retraction in cells expressing a PD-linked mutation in leucine-rich repeat kinase 2 (40). Whereas properly regulated autophagy plays a homeostatic and neuroprotective role, excessive or incomplete autophagy creates a condition of “autophagic stress” that can contribute to neurodegeneration (28).As mitochondrial fragmentation (3) and increased mitochondrial autophagy (4) have been described in human cells or tissues of PD patients, we investigated whether or not the engineered loss of PINK1 function could recapitulate these observations in human neuronal cells (SH-SY5Y). Stable knockdown of endogenous PINK1 gave rise to mitochondrial fragmentation and increased autophagy and mitophagy, whereas stable or transient overexpression of PINK1 had the opposite effect. Autophagy/mitophagy was dependent upon increased mitochondrial oxidant production and activation of fission. The data indicate that PINK1 is important for the maintenance of mitochondrial networks, suggesting that coordinated regulation of mitochondrial dynamics and autophagy limits cell death associated with loss of PINK1 function.  相似文献   

11.
12.
13.
Gluconeogenesis is an important metabolic pathway, which produces glucose from noncarbohydrate precursors such as organic acids, fatty acids, amino acids, or glycerol. Fructose-1,6-bisphosphatase, a key enzyme of gluconeogenesis, is found in all organisms, and five different classes of these enzymes have been identified. Here we demonstrate that Escherichia coli has two class II fructose-1,6-bisphosphatases, GlpX and YggF, which show different catalytic properties. We present the first crystal structure of a class II fructose-1,6-bisphosphatase (GlpX) determined in a free state and in the complex with a substrate (fructose 1,6-bisphosphate) or inhibitor (phosphate). The crystal structure of the ligand-free GlpX revealed a compact, globular shape with two α/β-sandwich domains. The core fold of GlpX is structurally similar to that of Li+-sensitive phosphatases implying that they have a common evolutionary origin and catalytic mechanism. The structure of the GlpX complex with fructose 1,6-bisphosphate revealed that the active site is located between two domains and accommodates several conserved residues coordinating two metal ions and the substrate. The third metal ion is bound to phosphate 6 of the substrate. Inorganic phosphate strongly inhibited activity of both GlpX and YggF, and the crystal structure of the GlpX complex with phosphate demonstrated that the inhibitor molecule binds to the active site. Alanine replacement mutagenesis of GlpX identified 12 conserved residues important for activity and suggested that Thr90 is the primary catalytic residue. Our data provide insight into the molecular mechanisms of the substrate specificity and catalysis of GlpX and other class II fructose-1,6-bisphosphatases.Fructose-1,6-bisphosphatase (FBPase,2 EC 3.1.3.11), a key enzyme of gluconeogenesis, catalyzes the hydrolysis of fructose 1,6-bisphosphate to form fructose 6-phosphate and orthophosphate. It is the reverse of the reaction catalyzed by phosphofructokinase in glycolysis, and the product, fructose 6-phosphate, is an important precursor in various biosynthetic pathways (1). In all organisms, gluconeogenesis is an important metabolic pathway that allows the cells to synthesize glucose from noncarbohydrate precursors, such as organic acids, amino acids, and glycerol. FBPases are members of the large superfamily of lithium-sensitive phosphatases, which includes three families of inositol phosphatases and FBPases (the phosphoesterase clan CL0171, 3167 sequences, Pfam data base). These enzymes show metal-dependent and lithium-sensitive phosphomonoesterase activity and include inositol polyphosphate 1-phosphatases, inositol monophosphatases (IMPases), 3′-phosphoadenosine 5′-phosphatases (PAPases), and enzymes acting on both inositol 1,4-bisphosphate and PAP (PIPases) (2). They possess a common structural core with the active site lying between α+β and α/β domains (3). Li+-sensitive phosphatases are putative targets for lithium therapy in the treatment of manic depressive patients (4), whereas FBPases are targets for the development of drugs for the treatment of noninsulin-dependent diabetes (5, 6). In addition, FBPase is required for virulence in Mycobacterium tuberculosis and Leishmania major and plays an important role in the production of lysine and glutamate by Corynebacterium glutamicum (7, 8).Presently, five different classes of FBPases have been proposed based on their amino acid sequences (FBPases I to V) (911). Eukaryotes contain only the FBPase I-type enzyme, but all five types exist in various prokaryotes. Types I, II, and III are primarily in bacteria, type IV in archaea (a bifunctional FBPase/inositol monophosphatase), and type V in thermophilic prokaryotes from both domains (11). Many organisms have more than one FBPase, mostly the combination of types I + II or II + III, but no bacterial genome has a combination of types I and III FBPases (9). The type I FBPase is the most widely distributed among living organisms and is the primary FBPase in Escherichia coli, most bacteria, a few archaea, and all eukaryotes (9, 1115). The type II FBPases are represented by the E. coli GlpX and FBPase F-I from Synechocystis PCC6803 (9, 16); type III is represented by the Bacillus subtilis FBPase (17); type IV is represented by the dual activity FBPases/inosine monophosphatases FbpA from Pyrococcus furiosus (18), MJ0109 from Methanococcus jannaschii (19), and AF2372 from Archaeoglobus fulgidus (20); and type V is represented by the FBPases TK2164 from Pyrococcus (Thermococcus) kodakaraensis and ST0318 from Sulfolobus tokodai (10, 21).Three-dimensional structures of the type I (from pig kidney, spinach chloroplasts, and E. coli), type IV (MJ0109 and AF2372), and type V (ST0318) FBPases have been solved (10, 11, 19, 20, 22, 23). FBPases I and IV and inositol monophosphatases share a common sugar phosphatase fold organized in five layered interleaved α-helices and β-sheets (α-β-α-β-α) (2, 19, 24). ST0318 (an FBPase V enzyme) is composed of one domain with a completely different four-layer α-β-β-α fold (10). The FBPases from these three classes (I, IV, and V) require divalent cations for activity (Mg2+, Mn2+, or Zn2+), and their structures have revealed the presence of three or four metal ions in the active site.E. coli has five Li+-sensitive phosphatases as follows: CysQ (a PAPase), SuhB (an IMPase), Fbp (a FBPase I enzyme), GlpX (a FBPase II), and YggF (an uncharacterized protein) (see the Pfam data base). CysQ is a 3′-phosphoadenosine 5′-phosphatase involved in the cysteine biosynthesis pathway (25, 26), whereas SuhB is an inositol monophosphatase (IMPase) that is also known as a suppressor of temperature-sensitive growth phenotypes in E. coli (27, 28). Fbp is required for growth on gluconeogenic substrates and probably represents the main gluconeogenic FBPase (12). This enzyme has been characterized both biochemically and structurally and shown to be inhibited by low concentrations of AMP (IC50 15 μm) (11, 29, 30). The E. coli GlpX, a class II enzyme FBPase, has been shown to possess a Mn2+-dependent FBPase activity (9). The increased expression of glpX from a multicopy plasmid complemented the Fbp- phenotype; however, the glpX knock-out strain grew normally on gluconeogenic substrates (succinate or glycerol) (9).In this study, we present the first structure of a class II FBPase, the E. coli GlpX, in a free state and in the complex with FBP + metals or phosphate. We have demonstrated that the fold of GlpX is similar to that of the lithium-sensitive phosphatases. We have identified the GlpX residues important for activity and proposed a catalytic mechanism. We have also showed that YggF is a third FBPase in E. coli, which has distinct catalytic properties and is more sensitive than GlpX to the inhibition by lithium or phosphate.  相似文献   

14.
We have investigated the possible biochemical basis for enhancements in NO production in endothelial cells that have been correlated with agonist- or shear stress-evoked phosphorylation at Ser-1179. We have found that a phosphomimetic substitution at Ser-1179 doubles maximal synthase activity, partially disinhibits cytochrome c reductase activity, and lowers the EC50(Ca2+) values for calmodulin binding and enzyme activation from the control values of 182 ± 2 and 422 ± 22 nm to 116 ± 2 and 300 ± 10 nm. These are similar to the effects of a phosphomimetic substitution at Ser-617 (Tran, Q. K., Leonard, J., Black, D. J., and Persechini, A. (2008) Biochemistry 47, 7557–7566). Although combining substitutions at Ser-617 and Ser-1179 has no additional effect on maximal synthase activity, cooperativity between the two substitutions completely disinhibits reductase activity and further reduces the EC50(Ca2+) values for calmodulin binding and enzyme activation to 77 ± 2 and 130 ± 5 nm. We have confirmed that specific Akt-catalyzed phosphorylation of Ser-617 and Ser-1179 and phosphomimetic substitutions at these positions have similar functional effects. Changes in the biochemical properties of eNOS produced by combined phosphorylation at Ser-617 and Ser-1179 are predicted to substantially increase synthase activity in cells at a typical basal free Ca2+ concentration of 50–100 nm.The nitric-oxide synthases catalyze formation of NO and l-citrulline from l-arginine and O2, with NADPH as the electron donor (1). The role of NO generated by endothelial nitricoxide synthase (eNOS)2 in the regulation of smooth muscle tone is well established and was the first of several physiological roles for this small molecule that have so far been identified (2). The nitric-oxide synthases are homodimers of 130–160-kDa subunits. Each subunit contains a reductase and oxygenase domain (1). A significant difference between the reductase domains in eNOS and nNOS and the homologous P450 reductases is the presence of inserts in these synthase isoforms that appear to maintain them in their inactive states (3, 4). A calmodulin (CaM)-binding domain is located in the linker that connects the reductase and oxygenase domains, and the endothelial and neuronal synthases both require Ca2+ and exogenous CaM for activity (5, 6). When CaM is bound, it somehow counteracts the effects of the autoinhibitory insert(s) in the reductase. The high resolution structure for the complex between (Ca2+)4-CaM and the isolated CaM-binding domain from eNOS indicates that the C-ter and N-ter lobes of CaM, which each contain a pair of Ca2+-binding sites, enfold the domain, as has been observed in several other such CaM-peptide complexes (7). Consistent with this structure, investigations of CaM-dependent activation of the neuronal synthase suggest that both CaM lobes must participate (8, 9).Bovine eNOS can be phosphorylated in endothelial cells at Ser-116, Thr-497, Ser-617, Ser-635, and Ser-1179 (1012). There are equivalent phosphorylation sites in the human enzyme (1012). Phosphorylation of the bovine enzyme at Thr-497, which is located in the CaM-binding domain, blocks CaM binding and enzyme activation (7, 11, 13, 14). Ser-116 can be basally phosphorylated in cells (10, 11, 13, 15), and dephosphorylation of this site has been correlated with increased NO production (13, 15). However, it has also been reported that a phosphomimetic substitution at this position has no effect on enzyme activity measured in vitro (13). Ser-1179 is phosphorylated in response to a variety of stimuli, and this has been reliably correlated with enhanced NO production in cells (10, 11). Indeed, NO production is elevated in transgenic endothelium expressing an eNOS mutant containing an S1179D substitution, but not in tissue expressing an S1179A mutant (16). Shear stress or insulin treatment is correlated with Akt-catalyzed phosphorylation of Ser-1179 in endothelial cells, and this is correlated with increased NO production in the absence of extracellular Ca2+ (1719). Akt-catalyzed phosphorylation or an S1179D substitution has also been correlated with increased synthase activity in cell extracts at low intracellular free [Ca2+] (17). Increased NO production has also been observed in cells expressing an eNOS mutant containing an S617D substitution, and physiological stimuli such as shear-stress, bradykinin, VEGF, and ATP appear to stimulate Akt-catalyzed phosphorylation of Ser-617 and Ser-1179 (12, 13, 20). Although S617D eNOS has been reported to have the same maximum activity in vitro as the wild type enzyme (20), in our hands an S617D substitution increases the maximal CaM-dependent synthase activity of purified mutant enzyme ∼2-fold, partially disinhibits reductase activity, and reduces the EC50(Ca2+) values for CaM binding and enzyme activation (21).In this report, we describe the effects of a phosphomimetic Asp substitution at Ser-1179 in eNOS on the Ca2+ dependence of CaM binding and CaM-dependent activation of reductase and synthase activities. We also describe the effects on these properties of combining this substitution with one at Ser-617. Finally, we demonstrate that Akt-catalyzed phosphorylation and Asp substitutions at Ser-617 and Ser-1179 have similar functional effects. Our results suggest that phosphorylation of eNOS at Ser-617 and Ser-1179 can substantially increase synthase activity in cells at a typical basal free Ca2+ concentration of 50–100 nm, while single phosphorylations at these sites produce smaller activity increases, and can do so only at higher free Ca2+ concentrations.  相似文献   

15.
Intersectin-short (intersectin-s) is a multimodule scaffolding protein functioning in constitutive and regulated forms of endocytosis in non-neuronal cells and in synaptic vesicle (SV) recycling at the neuromuscular junction of Drosophila and Caenorhabditis elegans. In vertebrates, alternative splicing generates a second isoform, intersectin-long (intersectin-l), that contains additional modular domains providing a guanine nucleotide exchange factor activity for Cdc42. In mammals, intersectin-s is expressed in multiple tissues and cells, including glia, but excluded from neurons, whereas intersectin-l is a neuron-specific isoform. Thus, intersectin-I may regulate multiple forms of endocytosis in mammalian neurons, including SV endocytosis. We now report, however, that intersectin-l is localized to somatodendritic regions of cultured hippocampal neurons, with some juxtanuclear accumulation, but is excluded from synaptophysin-labeled axon terminals. Consistently, intersectin-l knockdown (KD) does not affect SV recycling. Instead intersectin-l co-localizes with clathrin heavy chain and adaptor protein 2 in the somatodendritic region of neurons, and its KD reduces the rate of transferrin endocytosis. The protein also co-localizes with F-actin at dendritic spines, and intersectin-l KD disrupts spine maturation during development. Our data indicate that intersectin-l is indeed an important regulator of constitutive endocytosis and neuronal development but that it is not a prominent player in the regulated endocytosis of SVs.Clathrin-mediated endocytosis (CME)4 is a major mechanism by which cells take up nutrients, control the surface levels of multiple proteins, including ion channels and transporters, and regulate the coupling of signaling receptors to downstream signaling cascades (1-5). In neurons, CME takes on additional specialized roles; it is an important process regulating synaptic vesicle (SV) availability through endocytosis and recycling of SV membranes (6, 7), it shapes synaptic plasticity (8-10), and it is crucial in maintaining synaptic membranes and membrane structure (11).Numerous endocytic accessory proteins participate in CME, interacting with each other and with core components of the endocytic machinery such as clathrin heavy chain (CHC) and adaptor protein-2 (AP-2) through specific modules and peptide motifs (12). One such module is the Eps15 homology domain that binds to proteins bearing NPF motifs (13, 14). Another is the Src homology 3 (SH3) domain, which binds to proline-rich domains in protein partners (15). Intersectin is a multimodule scaffolding protein that interacts with a wide range of proteins, including several involved in CME (16). Intersectin has two N-terminal Eps15 homology domains that are responsible for binding to epsin, SCAMP1, and numb (17-19), a central coil-coiled domain that interacts with Eps15 and SNAP-23 and -25 (17, 20, 21), and five SH3 domains in its C-terminal region that interact with multiple proline-rich domain proteins, including synaptojanin, dynamin, N-WASP, CdGAP, and mSOS (16, 22-25). The rich binding capability of intersectin has linked it to various functions from CME (17, 26, 27) and signaling (22, 28, 29) to mitogenesis (30, 31) and regulation of the actin cytoskeleton (23).Intersectin functions in SV recycling at the neuromuscular junction of Drosophila and C. elegans where it acts as a scaffold, regulating the synaptic levels of endocytic accessory proteins (21, 32-34). In vertebrates, the intersectin gene is subject to alternative splicing, and a longer isoform (intersectin-l) is generated that is expressed exclusively in neurons (26, 28, 35, 36). This isoform has all the binding modules of its short (intersectin-s) counterpart but also has additional domains: a DH and a PH domain that provide guanine nucleotide exchange factor (GEF) activity specific for Cdc42 (23, 37) and a C2 domain at the C terminus. Through its GEF activity and binding to actin regulatory proteins, including N-WASP, intersectin-l has been implicated in actin regulation and the development of dendritic spines (19, 23, 24). In addition, because the rest of the binding modules are shared between intersectin-s and -l, it is generally thought that the two intersectin isoforms have the same endocytic functions. In particular, given the well defined role for the invertebrate orthologs of intersectin-s in SV endocytosis, it is thought that intersectin-l performs this role in mammalian neurons, which lack intersectin-s. Defining the complement of intersectin functional activities in mammalian neurons is particularly relevant given that the protein is involved in the pathophysiology of Down syndrome (DS). Specifically, the intersectin gene is localized on chromosome 21q22.2 and is overexpressed in DS brains (38). Interestingly, alterations in endosomal pathways are a hallmark of DS neurons and neurons from the partial trisomy 16 mouse, Ts65Dn, a model for DS (39, 40). Thus, an endocytic trafficking defect may contribute to the DS disease process.Here, the functional roles of intersectin-l were studied in cultured hippocampal neurons. We find that intersectin-l is localized to the somatodendritic regions of neurons, where it co-localizes with CHC and AP-2 and regulates the uptake of transferrin. Intersectin-l also co-localizes with actin at dendritic spines and disrupting intersectin-l function alters dendritic spine development. In contrast, intersectin-l is absent from presynaptic terminals and has little or no role in SV recycling.  相似文献   

16.
17.
PsbS plays a major role in activating the photoprotection mechanism known as “non-photochemical quenching,” which dissipates chlorophyll excited states exceeding the capacity for photosynthetic electron transport. PsbS activity is known to be triggered by low lumenal pH. However, the molecular mechanism by which this subunit regulates light harvesting efficiency is still unknown. Here we show that PsbS controls the association/dissociation of a five-subunit membrane complex, composed of two monomeric Lhcb proteins (CP29 and CP24) and the trimeric LHCII-M. Dissociation of this supercomplex is indispensable for the onset of non-photochemical fluorescence quenching in high light, strongly suggesting that protein subunits catalyzing the reaction of heat dissipation are buried into the complex and thus not available for interaction with PsbS. Consistently, we showed that knock-out mutants on two subunits participating to the B4C complex were strongly affected in heat dissipation. Direct observation by electron microscopy and image analysis showed that B4C dissociation leads to the redistribution of PSII within grana membranes. We interpreted these results to mean that the dissociation of B4C makes quenching sites, possibly CP29 and CP24, available for the switch to an energy-quenching conformation. These changes are reversible and do not require protein synthesis/degradation, thus allowing for changes in PSII antenna size and adaptation to rapidly changing environmental conditions.Photosynthetic reaction centers exploit solar energy to drive electrons from water to NADP+. This transport is coupled to H+ transfer from chloroplast stroma to thylakoids lumen, building a proton gradient for ATP synthesis (1). The capacity for light absorption is increased by pigment-binding proteins that compose the antenna system. In higher plants, the antenna is composed of the nuclear-encoded Chl4 a/b-binding light-harvesting complexes (Lhc). The major constituent of the photosystem II (PSII) outer antenna is LHCII, a heterotrimer composed by different combinations of Lhcb1, Lhcb2, and Lhcb3 gene products (2). Three additional monomeric antenna complexes (CP29, CP26, and CP24) encoded by the lhcb4, lhcb5, and lhcb6 genes, respectively, are localized in between the core complex and LHCII (3). Similarly, PSI has four Lhca antenna proteins, yielding a total of 10 distinct Lhc isoforms (2). Differences between the mentioned isoforms have been largely conserved in all higher plants during at the last 350 million years of evolution, strongly indicating that each pigment-protein complex has a specific function (4), although the specific role of each gene product in light harvesting and/or photoprotection is still under debate (5). Their topological organization into the supercomplex has been analyzed by electron microscopy and biochemical methods (3, 6-8) showing that Lhcb subunits are organized into two layers around the PSII core. The inner layer is composed of CP29, CP26, and the S-type LHCII trimer, forming, together with the PSII core, the so-called C2S2 particle (9, 10). The outer layer is made of LHCII trimers and CP24, to build up the larger C2S2M2Lx complexes (9) in which the number of LHCII-L trimers depends on the light intensity during growth (11, 12).This structural organization responds to the requirements of light harvesting regulation; in high light, when absorbed energy does not limit growth, the PSII antenna loses the components of the external antenna layer, namely CP24, LHCII-M, and LHCII-L (12), whereas the internal antenna components, CP26, CP29, and LHCII-S, are always retained in a 1:1 stoichiometry with the PSII core complex. This is consistent with the composition of a mutant exhibiting chronic plastoquinone reduction, mimicking overexcitation (10). Such an acclimation to contrasting light conditions, however, requires days to weeks (12, 13), whereas plants are often exposed to rapid changes in light intensity, temperature, and water availability. In these conditions, incomplete photochemical quenching leads to an increased Chl excited state (1Chl*) lifetime and increased probability of Chl a triplet formation (3Chl*) by intersystem crossing. Chl triplets react with oxygen (3O2) and form harmful reactive oxygen species, responsible for photoinhibition and oxidative stress (14). These harmful events are counteracted by photoprotection mechanisms consisting either in the scavenging of generated reactive oxygen species (15) or in prevention of their production through dissipation of the 1Chl* in excess (16, 17). This latter process is known as non-photochemical quenching (NPQ) and is observed as light-dependent quenching of Chl fluorescence. NPQ has been shown to be composed by at least two components with different activation time scales. The first, feedback de-excitation quenching (qE), is rapidly activated upon increasing light intensity, whereas a second component (qI) is slower. Full NPQ activation requires zeaxanthin synthesis (18, 19) and the PsbS protein (20) that senses low lumenal pH through two lumen-exposed protonatable residues (21, 22). Mutants lacking Chl b, and thus lacking Lhc proteins, or exhibiting alterations in the topological organization of PSII antenna also undergo a strong reduction in NPQ, demonstrating the involvement of antenna proteins in its activation (23-25).In this work we analyzed the changes in the organization of the PSII antenna during exposure to strong light and NPQ development. We found that a supramolecular complex, named B4C, composed of CP29, CP24, and LHCII-M, connects the inner and outer antenna moieties, dissociates during light exposure, and reassociates during subsequent dark recovery. Dissociation of the B4C complex appears necessary for the establishment of non-photochemical fluorescence quenching. Upon illumination, PSII distribution within grana membranes was also affected, and we observed a shorter distance between PSII reaction centers, implying enrichment in C2S2 complexes and depletion in the outer LHCII components. These results suggest that the NPQ process includes a rapid and reversible change in the organization of grana membranes with disconnection of a subset of Lhcb proteins from the PSII reaction center.  相似文献   

18.
Zinc is hypothesized to be co-released with glutamate at synapses of the central nervous system. Zinc binds to NR1/NR2A N-methyl-d-aspartate (NMDA) receptors with high affinity and inhibits NMDAR function in a voltage-independent manner. The serine protease plasmin can cleave a number of substrates, including protease-activated receptors, and may play an important role in several disorders of the central nervous system, including ischemia and spinal cord injury. Here, we demonstrate that plasmin can cleave the native NR2A amino-terminal domain (NR2AATD), removing the functional high affinity Zn2+ binding site. Plasmin also cleaves recombinant NR2AATD at lysine 317 (Lys317), thereby producing a ∼40-kDa fragment, consistent with plasmin-induced NR2A cleavage fragments observed in rat brain membrane preparations. A homology model of the NR2AATD predicts that Lys317 is near the surface of the protein and is accessible to plasmin. Recombinant expression of NR2A with an amino-terminal deletion at Lys317 is functional and Zn2+ insensitive. Whole cell voltage-clamp recordings show that Zn2+ inhibition of agonist-evoked NMDA receptor currents of NR1/NR2A-transfected HEK 293 cells and cultured cortical neurons is significantly reduced by plasmin treatment. Mutating the plasmin cleavage site Lys317 on NR2A to alanine blocks the effect of plasmin on Zn2+ inhibition. The relief of Zn2+ inhibition by plasmin occurs in PAR1-/- cortical neurons and thus is independent of interaction with protease-activated receptors. These results suggest that plasmin can directly interact with NMDA receptors, and plasmin may increase NMDA receptor responses through disruption or removal of the amino-terminal domain and relief of Zn2+ inhibition.N-Methyl-d-aspartate (NMDA)2 receptors are one of three types of ionotropic glutamate receptors that play critical roles in excitatory neurotransmission, synaptic plasticity, and neuronal death (13). NMDA receptors are comprised of glycine-binding NR1 subunits in combination with at least one type of glutamate-binding NR2 subunit (1, 4). Each subunit contains three transmembrane domains, one cytoplasmic re-entrant membrane loop, one bi-lobed domain that forms the ligand binding site, and one bi-lobed amino-terminal domain (ATD), thought to share structural homology to periplasmic amino acid-binding proteins (46). Activation of NMDA receptors requires combined stimulation by glutamate and the co-agonist glycine in addition to membrane depolarization to overcome voltage-dependent Mg2+ block of the ion channel (7). The activity of NMDA receptors is negatively modulated by a variety of extracellular ions, including Mg2+, polyamines, protons, and Zn2+ ions, which can exert tonic inhibition under physiological conditions (1, 4). Several extracellular modulators such as Zn2+ and ifenprodil are thought to act at the ATD of the NMDA receptor (814).Zinc is a transition metal that plays key roles in both catalytic and structural capacities in all mammalian cells (15). Zinc is required for normal growth and survival of cells. In addition, neuronal death in hypoxia-ischemia and epilepsy has been associated with Zn2+ (1618). Abnormal metabolism of zinc may contribute to induction of cytotoxicity in neurodegenerative diseases, such as Alzheimer''s disease, Parkinson''s disease, and amyotrophic lateral sclerosis (19). Zinc is co-released with glutamate at excitatory presynaptic terminals and inhibits native NMDA receptor activation (20, 21). Zn2+ inhibits NMDA receptor function through a dual mechanism, which includes voltage-dependent block and voltage-independent inhibition (2224). Voltage-independent Zn2+ inhibition at low nanomolar concentrations (IC50, 20 nm) is observed for NR2A-containing NMDA receptors (2528). Evidence has accumulated that the amino-terminal domain of the NR2A subunit controls high-affinity Zn2+ inhibition of NMDA receptors, and several histidine residues in this region may constitute part of an NR2A-specific Zn2+ binding site (8, 9, 11, 12). For the NR2A subunit, several lines of evidence suggest that Zn2+ acts by enhancing proton inhibition (8, 11, 29, 30).Serine proteases present in the circulation, mast cells, and elsewhere signal directly to cells by cleaving protease-activated receptors (PARs), members of a subfamily of G-protein-coupled receptors. Cleavage exposes a tethered ligand domain that binds to and activates the cleaved receptors (31, 32). Protease receptor activation has been studied extensively in relation to coagulation and thrombolysis (33). In addition to their circulation in the bloodstream, some serine proteases and PARs are expressed in the central nervous system, and have been suggested to play roles in physiological conditions (e.g. long-term potentiation or memory) and pathophysiological states such as glial scarring, edema, seizure, and neuronal death (31, 3436).Functional interactions between proteases and NMDA receptors have previously been suggested. Earlier studies reported that the blood-derived serine protease thrombin potentiates NMDA receptor response more than 2-fold through activation of PAR1 (37). Plasmin, another serine protease, similarly potentiates NMDA receptor response (38). Tissue-plasminogen activator (tPA), which catalyzes the conversion of the zymogen precursor plasminogen to plasmin and results in PAR1 activation, also interacts with and cleaves the ATD of the NR1 subunit of the NMDA receptor (39, 40). This raises the possibility that plasmin may also interact directly with the NMDA receptor subunits to modulate receptor response. We therefore investigated the ability of plasmin to cleave the NR2A NMDA receptor subunit. We found that nanomolar concentrations of plasmin can cleave within the ATD, a region that mediates tonic voltage-independent Zn2+ inhibition of NR2A-containing NMDA receptors. We hypothesized that plasmin cleavage reduces the Zn2+-mediated inhibition of NMDA receptors by removing the Zn2+ binding domain. In the present study, we have demonstrated that Zn2+ inhibition of agonist-evoked NMDA currents is decreased significantly by plasmin treatment in recombinant NR1/NR2A-transfected HEK 293 cells and cultured cortical neurons. These concentrations of plasmin may be pathophysiologically relevant in situations in which the blood-brain barrier is compromised, which could allow blood-derived plasmin to enter brain parenchyma at concentrations in excess of these that can cleave NR2A. Thus, ability of plasmin to potentiate NMDA function through the relief of the Zn2+ inhibition could exacerbate the harmful actions of NMDA receptor overactivation in pathological situations. In addition, if newly cleaved NR2AATD enters the bloodstream during ischemic injury, it could serve as a biomarker of central nervous system injury.  相似文献   

19.
Melanoma cells express the chemokine receptor CXCR4 that confers high invasiveness upon binding to its ligand CXCL12. Melanoma cells at initial stages of the disease show reduction or loss of E-cadherin expression, but recovery of its expression is frequently found at advanced phases. We overexpressed E-cadherin in the highly invasive BRO lung metastatic cell melanoma cell line to investigate whether it could influence CXCL12-promoted cell invasion. Overexpression of E-cadherin led to defective invasion of melanoma cells across Matrigel and type I collagen in response to CXCL12. A decrease in individual cell migration directionality toward the chemokine and reduced adhesion accounted for the impaired invasion. A p190RhoGAP-dependent inhibition of RhoA activation was responsible for the impairment in chemokine-stimulated E-cadherin melanoma transfectant invasion. Furthermore, we show that p190RhoGAP and p120ctn associated predominantly on the plasma membrane of cells overexpressing E-cadherin, and that E-cadherin-bound p120ctn contributed to RhoA inactivation by favoring p190RhoGAP-RhoA association. These results suggest that melanoma cells at advanced stages of the disease could have reduced metastatic potency in response to chemotactic stimuli compared with cells lacking E-cadherin, and the results indicate that p190RhoGAP is a central molecule controlling melanoma cell invasion.Cadherins are a family of Ca2+-dependent adhesion molecules that mediate cell-cell contacts and are expressed in most solid tissues providing a tight control of morphogenesis (1, 2). Classical cadherins, such as epithelial (E) cadherin, are found in adherens junctions, forming core protein complexes with β-catenin, α-catenin, and p120 catenin (p120ctn). Both β-catenin and p120ctn directly interact with E-cadherin, whereas α-catenin associates with the complex through its binding to β-catenin, providing a link with the actin cytoskeleton (1, 2). E-cadherin is frequently lost or down-regulated in many human tumors, coincident with morphological epithelial to mesenchymal transition and acquisition of invasiveness (3-6).Although melanoma only accounts for 5% of skin cancers, when metastasis starts, it is responsible for 80% of deaths from skin cancers (7). Melanocytes express E-cadherin (8-10), but melanoma cells at early radial growth phase show a large reduction in the expression of this cadherin, and surprisingly, expression has been reported to be partially recovered by vertical growth phase and metastatic melanoma cells (9, 11, 12).Trafficking of cancer cells from primary tumor sites to intravasation into blood circulation and later to extravasation to colonize distant organs requires tightly regulated directional cues and cell migration and invasion that are mediated by chemokines, growth factors, and adhesion molecules (13). Solid tumor cells express chemokine receptors that provide guidance of these cells to organs where their chemokine ligands are expressed, constituting a homing model resembling the one used by immune cells to exert their immune surveillance functions (14). Most solid cancer cells express CXCR4, a receptor for the chemokine CXCL12 (also called SDF-1), which is expressed in lungs, bone marrow, and liver (15). Expression of CXCR4 in human melanoma has been detected in the vertical growth phase and on regional lymph nodes, which correlated with poor prognosis and increased mortality (16, 17). Previous in vivo experiments have provided evidence supporting a crucial role for CXCR4 in the metastasis of melanoma cells (18).Rho GTPases control the dynamics of the actin cytoskeleton during cell migration (19, 20). The activity of Rho GTPases is tightly regulated by guanine-nucleotide exchange factors (GEFs),4 which stimulate exchange of bound GDP by GTP, and inhibited by GTPase-activating proteins (GAPs), which promote GTP hydrolysis (21, 22), whereas guanine nucleotide dissociation inhibitors (GDIs) appear to mediate blocking of spontaneous activation (23). Therefore, cell migration is finely regulated by the balance between GEF, GAP, and GDI activities on Rho GTPases. Involvement of Rho GTPases in cancer is well documented (reviewed in Ref. 24), providing control of both cell migration and growth. RhoA and RhoC are highly expressed in colon, breast, and lung carcinoma (25, 26), whereas overexpression of RhoC in melanoma leads to enhancement of cell metastasis (27). CXCL12 activates both RhoA and Rac1 in melanoma cells, and both GTPases play key roles during invasion toward this chemokine (28, 29).Given the importance of the CXCL12-CXCR4 axis in melanoma cell invasion and metastasis, in this study we have addressed the question of whether changes in E-cadherin expression on melanoma cells might affect cell invasiveness. We show here that overexpression of E-cadherin leads to impaired melanoma cell invasion to CXCL12, and we provide mechanistic characterization accounting for the decrease in invasion.  相似文献   

20.
Cytokinesis in bacteria depends upon the contractile Z ring, which is composed of dynamic polymers of the tubulin homolog FtsZ as well as other membrane-associated proteins such as FtsA, a homolog of actin that is required for membrane attachment of the Z ring and its subsequent constriction. Here we show that a previously characterized hypermorphic mutant FtsA (FtsA*) partially disassembled FtsZ polymers in vitro. This effect was strictly dependent on ATP or ADP binding to FtsA* and occurred at substoichiometric levels relative to FtsZ, similar to cellular levels. Nucleotide-bound FtsA* did not affect FtsZ GTPase activity or the critical concentration for FtsZ assembly but was able to disassemble preformed FtsZ polymers, suggesting that FtsA* acts on FtsZ polymers. Microscopic examination of the inhibited FtsZ polymers revealed a transition from long, straight polymers and polymer bundles to mainly short, curved protofilaments. These results indicate that a bacterial actin, when activated by adenine nucleotides, can modify the length distribution of bacterial tubulin polymers, analogous to the effects of actin-depolymerizing factor/cofilin on F-actin.Bacterial cell division requires a large number of proteins that colocalize to form a putative protein machine at the cell membrane (1). This machine, sometimes called the divisome, recruits enzymes to synthesize the septum cell wall and to initiate and coordinate the invagination of the cytoplasmic membrane (and in Gram-negative bacteria, the outer membrane). The most widely conserved and key protein for this process is FtsZ, a homolog of tubulin that forms a ring structure called the Z ring, which marks the site of septum formation (2, 3). Like tubulin, FtsZ assembles into filaments with GTP but does not form microtubules (4). The precise assembly state and conformation of these FtsZ filaments at the division ring is not clear, although recent electron tomography work suggests that the FtsZ ring consists of multiple short filaments tethered to the membrane at discrete junctures (5), which may represent points along the filaments bridged by membrane anchor proteins.In Escherichia coli, two of these anchor proteins are known. One of these, ZipA, is not well conserved but is an essential protein in E. coli. ZipA binds to the C-terminal tail of FtsZ (68), and purified ZipA promotes bundling of FtsZ filaments in vitro (9, 10). The other, FtsA, is also essential in E. coli and is more widely conserved among bacterial species. FtsA is a member of the HSP70/actin superfamily (11, 12), and like ZipA, it interacts with the C-terminal tail of FtsZ (7, 1315). FtsA can self-associate (16, 17) and bind ATP (12, 18), but reports of ATPase activity vary, with Bacillus subtilis FtsA having high activity (19) and Streptococcus pneumoniae FtsA exhibiting no detectable activity (20). There are no reports of any other in vitro activities of FtsA, including effects on FtsZ assembly.Understanding how FtsA affects FtsZ assembly is important because FtsA has a number of key activities in the cell. It is required for recruitment of a number of divisome proteins (21, 22) and helps to tether the Z ring to the membrane via a C-terminal membrane-targeting sequence (23). FtsA, like ZipA and other divisome proteins, is necessary to activate the contraction of the Z ring (24, 25). In E. coli, the FtsA:FtsZ ratio is crucial for proper cell division, with either too high or too low a ratio inhibiting septum formation (26, 27). This ratio is roughly 1:5, with ∼700 molecules of FtsA and 3200 molecules of FtsZ per cell (28), which works out to concentrations of 1–2 and 5–10 μm, respectively.Another interesting property of FtsA is that single residue alterations in the protein can result in significant enhancement of divisome activity. For example, the R286W mutation of FtsA, also called FtsA*, can substitute for the native FtsA and divide the cell. However, this mutant FtsA causes E. coli cells to divide at less than 80% of their normal length (29) and allows efficient division of E. coli cells in the absence of ZipA (30), indicating that it has gain-of-function activity. FtsA* and other hypermorphic mutations such as E124A and I143L can also increase division activity in cells lacking other essential divisome components (3133). The R286W and E124A mutants of FtsA also bypass the FtsA:FtsZ ratio rule, allowing cell division to occur at higher ratios than with WT2 FtsA. This may be because the altered FtsA proteins self-associate more readily than WT FtsA, which may cause different changes in FtsZ assembly state as compared with WT FtsA (17, 34).In this study, we use an in vitro system with purified FtsZ and a purified tagged version of FtsA* to elucidate the role of FtsA in activating constriction of the Z ring in vivo. We show that FtsA*, at physiological concentrations in the presence of ATP or ADP, has significant effects on the assembly of FtsZ filaments.  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