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 62 毫秒
1.
Recent studies have ascribed many non-pumping functions to the Na/K-ATPase. We show here that graded knockdown of cellular Na/K-ATPase α1 subunit produces a parallel decrease in both caveolin-1 and cholesterol in light fractions of LLC-PK1 cell lysates. This observation is further substantiated by imaging analyses, showing redistribution of cholesterol from the plasma membrane to intracellular compartments in the knockdown cells. Moreover, this regulation is confirmed in α1+/– mouse liver. Functionally, the knockdown-induced redistribution appears to affect the cholesterol sensing in the endoplasmic reticulum, because it activates the sterol regulatory element-binding protein pathway and increases expression of hydroxymethylglutaryl-CoA reductase and low density lipoprotein receptor in the liver. Consistently, we detect a modest increase in hepatic cholesterol as well as a reduction in the plasma cholesterol. Mechanistically, α1+/– livers show increases in cellular Src and ERK activity and redistribution of caveolin-1. Although activation of Src is not required in Na/K-ATPase-mediated regulation of cholesterol distribution, the interaction between the Na/K-ATPase and caveolin-1 is important for this regulation. Taken together, our new findings demonstrate a novel function of the Na/K-ATPase in control of the plasma membrane cholesterol distribution. Moreover, the data also suggest that the plasma membrane Na/K-ATPase-caveolin-1 interaction may represent an important sensing mechanism by which the cells regulate the sterol regulatory element-binding protein pathway.The Na/K-ATPase, also called the sodium pump, is an ion transporter that mediates active transport of Na+ and K+ across the plasma membrane by hydrolyzing ATP (1, 2). The functional sodium pump is mainly composed of α and β subunits. The α subunit is the catalytic component of the holoenzyme; it contains both the nucleotide and the cation binding sites (3). So far, four isoforms of α subunit have been discovered, and each one shows a distinct tissue distribution pattern (4, 5). Interestingly, studies during the past few years have uncovered many non-pumping functions of Na/K-ATPase (610). Recently, we have demonstrated that more than half of the Na/K-ATPase may actually perform cellular functions other than ion pumping at least in LLC-PK1 cells (11). Moreover, the non-pumping pool of Na/K-ATPase mainly resides in caveolae and interacts with a variety of proteins such as Src, inositol 1,4,5-trisphosphate receptor, and caveolin-1 (1214). While the interaction between Na/K-ATPase and inositol 1,4,5-trisphosphate receptor facilitates Ca2+ signaling (13) the dynamic association between Na/K-ATPase and Src appears to be an essential step for ouabain to stimulate cellular kinases (15). More recently, we report that the interaction between the Na/K-ATPase and caveolin-1 plays an important role for the membrane trafficking of caveolin-1. Knockdown of the Na/K-ATPase leads to altered subcellular distribution of caveolin-1 and increases the mobility of caveolin-1-containing vesicles (16).Caveolin is a protein marker for caveolae (17). Caveolae are flask-shaped vesicular invaginations of plasma membrane and are enriched in cholesterol, glycosphingolipids, and sphingomyelin (18). There are three genes and six isoforms of caveolin. Caveolin-1 is a 22-kDa protein and is expressed in many types of cells, including epithelial and endothelial cells. In addition to their role in biogenesis of caveolae (19), accumulating evidence has implicated caveolin proteins in cellular cholesterol homeostasis (20). For instance, caveolin-1 directly binds to cholesterol in a 1:1 ratio (21). It was also found to be an integral member of the intracellular cholesterol trafficking machinery between internal membranes and plasma membrane (22, 23). The expression of caveolin-1 appears to be under control of SREBPs,2 the master regulators of intracellular cholesterol level (24). Furthermore, knockout of caveolin-1 significantly affected cholesterol metabolism in mouse embryonic fibroblasts and mouse peritoneal macrophages (25). Because we found that the Na/K-ATPase regulates cellular distribution of caveolin-1, we propose that it may also affect intracellular cholesterol distribution and metabolism. To test our hypothesis, we have investigated whether sodium pump α1 knockdown affects cholesterol distribution and metabolism both in vitro and in vivo. Our results indicate that sodium pump α1 expression level plays a role in the proper distribution of intracellular cholesterol. Down-regulation of sodium pump α1 not only redistributes cholesterol between the plasma membrane and cytosolic compartments, but also alters cholesterol metabolism in mice.  相似文献   

2.
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   

3.
4.
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   

5.
6.
7.
It has recently been shown that the epithelial Na+ channel (ENaC) is compartmentalized in caveolin-rich lipid rafts and that pharmacological depletion of membrane cholesterol, which disrupts lipid raft formation, decreases the activity of ENaC. Here we show, for the first time, that a signature protein of caveolae, caveolin-1 (Cav-1), down-regulates the activity and membrane surface expression of ENaC. Physical interaction between ENaC and Cav-1 was also confirmed in a coimmunoprecipitation assay. We found that the effect of Cav-1 on ENaC requires the activity of Nedd4-2, a ubiquitin protein ligase of the Nedd4 family, which is known to induce ubiquitination and internalization of ENaC. The effect of Cav-1 on ENaC requires the proline-rich motifs at the C termini of the β- and γ-subunits of ENaC, the binding motifs that mediate interaction with Nedd4-2. Taken together, our data suggest that Cav-1 inhibits the activity of ENaC by decreasing expression of ENaC at the cell membrane via a mechanism that involves the promotion of Nedd4-2-dependent internalization of the channel.Amiloride-sensitive epithelial Na+ channels (ENaC)3 are membrane proteins that are expressed in salt-absorptive epithelia, including the distal collecting tubules of the kidney, the mucosa of the distal colon, the respiratory epithelium, and the excretory ducts of sweat and salivary glands (14). Na+ absorption via ENaC is critical to the normal regulation of Na+ and fluid homeostasis and is important for maintaining blood pressure (5) and the volume of fluid in the respiratory passages (6). Increased ENaC activity has been implicated in the salt-sensitive inherited form of hypertension, Liddle''s syndrome (7), and dehydration of the surface of the airway epithelium in the pathology associated with cystic fibrosis lung disease (8).Expression of ENaC at the cell membrane surface is regulated by the E3 ubiquitin protein ligase, Nedd4-2 (neural precursor cell expressed developmentally down-regulated protein 4) (9). Interaction between the WW domains of Nedd4-2 and the proline-rich PY motifs (PPPXY) on ENaC is essential for Nedd4-2 to exert a negative effect on the channel (10, 11). This interaction leads to ubiquitination-dependent internalization of ENaC (12, 13). Several regulators of ENaC exert their effects on the channel by modulating the action of Nedd4-2. For instance, serum and glucocorticoid-dependent protein kinase (14), protein kinase B (15), and G protein-coupled receptor kinase (16) up-regulate activity of ENaC by inhibiting Nedd4-2. Although the details of cellular mechanisms that underlie internalization of ENaC remain to be elucidated, the physiological significance of Nedd4-dependent internalization of the channel has been well established. For instance, heritable mutations that delete the cytosolic termini of the β-or γ-subunit of ENaC, which contain the proline-rich motifs, are known to cause hyperactivity of ENaC in the kidney (17) and increase cell surface expression of the channel (7, 18).The plasma membranes of most cell types contain lipid raft microdomains that are enriched with glycosphingolipid and cholesterol (19), that have distinctive biophysical properties, and that selectively include or exclude signaling molecules (20). These microdomains promote clustering of an array of integral membrane proteins in the membrane leaflets (21) and may be important for organizing cascades of signaling molecules (22, 23). Processes in which raft microdomains are involved include the intracellular transport of proteins and lipids to the cell membrane (24), the endocytotic retrieval of membrane proteins (25, 26), and signal transduction (27, 28). In addition, segregation of signaling molecules within lipid rafts may facilitate cross-talk between signal transduction pathways (29), a phenomenon that may be important in ensuring rapid and efficient integration of multiple cellular signaling events (30, 31). Of particular interest is the subpopulation of lipid rafts enriched with caveolin proteins. Caveolin-1 (Cav-1), a major caveolin isoform expressed in nonmuscle cells, has been identified as being involved in diverse cellular functions, such as vesicular transport, cholesterol homeostasis, and signal transduction (32). Cav-1 also regulates the activity and membrane expression of ion channels and transporters (28).In epithelia, the majority of lipid rafts exist at the apical membrane surface (22). Pools of ENaC (3336) and several proteins that regulate activity of ENaC, such as Nedd4 (37), protein kinase B (38), protein kinase C (39), Go (40), and the G protein-coupled receptor kinase (41), have been identified in detergent-insoluble and cholesterol-rich membrane fractions from a variety of cell types, consistent with localization of these proteins in lipid rafts. Furthermore, detergent-free buoyant density separation of lipid rafts has revealed the presence of Cav-1 with ENaC in the lipid raft-rich membrane fraction (35). The physiological role of lipid rafts in the regulation of ENaC has been the subject of many recent investigations. Most of these studies used a pharmacological agent, methyl-β-cyclodextrin (MβCD), to promote redistribution of proteins away from the cholesterol-enriched membrane domains. The results were, however, inconclusive. In some studies, MβCD treatment was found to inhibit open probability (42) or cell surface expression of ENaC (35), whereas others found no direct effect of MβCD on the channel (33, 43).Despite a number of studies into the role of lipid rafts on the regulation of ENaC, little is known about the physiological relevance of caveolins to the function of this ion channel. In the present study, we use gene interference and gene expression techniques to determine the role of Cav-1 in the regulation of ENaC activity. We provide evidence of the association of Cav-1 with ENaC and evidence that Cav-1 negatively regulates both activity and abundance of ENaC at the surface of epithelial cells. Importantly, we demonstrate, for the first time, that the mechanism by which Cav-1 regulates activity of ENaC involves the E3 ubiquitin protein ligase, Nedd4-2.  相似文献   

8.
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   

9.
As obligate intracellular parasites, viruses exploit diverse cellular signaling machineries, including the mitogen-activated protein-kinase pathway, during their infections. We have demonstrated previously that the open reading frame 45 (ORF45) of Kaposi sarcoma-associated herpesvirus interacts with p90 ribosomal S6 kinases (RSKs) and strongly stimulates their kinase activities (Kuang, E., Tang, Q., Maul, G. G., and Zhu, F. (2008) J. Virol. 82 ,1838 -1850). Here, we define the mechanism by which ORF45 activates RSKs. We demonstrated that binding of ORF45 to RSK increases the association of extracellular signal-regulated kinase (ERK) with RSK, such that ORF45, RSK, and ERK formed high molecular mass protein complexes. We further demonstrated that the complexes shielded active pERK and pRSK from dephosphorylation. As a result, the complex-associated RSK and ERK were activated and sustained at high levels. Finally, we provide evidence that this mechanism contributes to the sustained activation of ERK and RSK in Kaposi sarcoma-associated herpesvirus lytic replication.The extracellular signal-regulated kinase (ERK)2 mitogen-activated protein kinase (MAPK) signaling pathway has been implicated in diverse cellular physiological processes including proliferation, survival, growth, differentiation, and motility (1-4) and is also exploited by a variety of viruses such as Kaposi sarcoma-associated herpesvirus (KSHV), human cytomegalovirus, human immunodeficiency virus, respiratory syncytial virus, hepatitis B virus, coxsackie, vaccinia, coronavirus, and influenza virus (5-17). The MAPK kinases relay the extracellular signaling through sequential phosphorylation to an array of cytoplasmic and nuclear substrates to elicit specific responses (1, 2, 18). Phosphorylation of MAPK is reversible. The kinetics of deactivation or duration of signaling dictates diverse biological outcomes (19, 20). For example, sustained but not transient activation of ERK signaling induces the differentiation of PC12 cells into sympathetic-like neurons and transformation of NIH3T3 cells (20-22). During viral infection, a unique biphasic ERK activation has been observed for some viruses (an early transient activation triggered by viral binding or entry and a late sustained activation correlated with viral gene expression), but the responsible viral factors and underlying mechanism for the sustained ERK activation remain largely unknown (5, 8, 13, 23).The p90 ribosomal S6 kinases (RSKs) are a family of serine/threonine kinases that lie at the terminus of the ERK pathway (1, 24-26). In mammals, four isoforms are known, RSK1 to RSK4. Each one has two catalytically functional kinase domains, the N-terminal kinase domain (NTKD) and C-terminal kinase domain (CTKD) as well as a linker region between the two. The NTKD is responsible for phosphorylation of exogenous substrates, and the CTKD and linker region regulate RSK activation (1, 24, 25). In quiescent cells ERK binds to the docking site in the C terminus of RSK (27-29). Upon mitogen stimulation, ERK is activated by its upstream MAPK/ERK kinase (MEK). The active ERK phosphorylates Thr-359/Ser-363 of RSK in the linker region (amino acid numbers refer to human RSK1) and Thr-573 in the CTKD activation loop. The activated CTKD then phosphorylates Ser-380 in the linker region, creating a docking site for 3-phosphoinositide-dependent protein kinase-1. The 3-phosphoinositide-dependent protein kinase-1 phosphorylates Ser-221 of RSK in the activation loop and activates the NTKD. The activated NTKD autophosphorylates the serine residue near the ERK docking site, causing a transient dissociation of active ERK from RSK (25, 26, 28). The stimulation of quiescent cells by a mitogen such as epidermal growth factor or a phorbol ester such as 12-O-tetradecanoylphorbol-13-acetate (TPA) usually results in a transient RSK activation that lasts less than 30 min. RSKs have been implicated in regulating cell survival, growth, and proliferation. Mutation or aberrant expression of RSK has been implicated in several human diseases including Coffin-Lowry syndrome and prostate and breast cancers (1, 24, 25, 30-32).KSHV is a human DNA tumor virus etiologically linked to Kaposi sarcoma, primary effusion lymphoma, and a subset of multicentric Castleman disease (33, 34). Infection and reactivation of KSHV activate multiple MAPK pathways (6, 12, 35). Noticeably, the ERK/RSK activation is sustained late during KSHV primary infection and reactivation from latency (5, 6, 12, 23), but the mechanism of the sustained ERK/RSK activation is unclear. Recently, we demonstrated that ORF45, an immediate early and also virion tegument protein of KSHV, interacts with RSK1 and RSK2 and strongly stimulates their kinase activities (23). We also demonstrated that the activation of RSK plays an essential role in KSHV lytic replication (23). In the present study we determined the mechanism of ORF45-induced sustained ERK/RSK activation. We found that ORF45 increases the association of RSK with ERK and protects them from dephosphorylation, causing sustained activation of both ERK and RSK.  相似文献   

10.
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   

11.
12.
13.
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   

14.
Macroautophagy is a vacuolar lysosomal catabolic pathway that is stimulated during periods of nutrient starvation to preserve cell integrity. Ceramide is a bioactive sphingolipid associated with a large range of cell processes. Here we show that short-chain ceramides (C2-ceramide and C6-ceramide) and stimulation of the de novo ceramide synthesis by tamoxifen induce the dissociation of the complex formed between the autophagy protein Beclin 1 and the anti-apoptotic protein Bcl-2. This dissociation is required for macroautophagy to be induced either in response to ceramide or to starvation. Three potential phosphorylation sites, Thr69, Ser70, and Ser87, located in the non-structural N-terminal loop of Bcl-2, play major roles in the dissociation of Bcl-2 from Beclin 1. We further show that activation of c-Jun N-terminal protein kinase 1 by ceramide is required both to phosphorylate Bcl-2 and to stimulate macroautophagy. These findings reveal a new aspect of sphingolipid signaling in up-regulating a major cell process involved in cell adaptation to stress.Macroautophagy (referred to below as “autophagy”) is a vacuolar, lysosomal degradation pathway for cytoplasmic constituents that is conserved in eukaryotic cells (13). Autophagy is initiated by the formation of a multimembrane-bound autophagosome that engulfs cytoplasmic proteins and organelles. The last stage in the process results in fusion with the lysosomal compartments, where the autophagic cargo undergoes degradation. Basal autophagy is important in controlling the quality of the cytoplasm by removing damaged organelles and protein aggregates. Inhibition of basal autophagy in the brain is deleterious, and leads to neurodegeneration in mouse models (4, 5). Stimulation of autophagy during periods of nutrient starvation is a physiological response present at birth and has been shown to provide energy in various tissues of newborn pups (6). In cultured cells, starvation-induced autophagy is an autonomous cell survival mechanism, which provides nutrients to maintain a metabolic rate and level of ATP compatible with cell survival (7). In addition, starvation-induced autophagy blocks the induction of apoptosis (8). In other contexts, such as drug treatment and a hypoxic environment, autophagy has also been shown to be cytoprotective in cancer cells (9, 10). However, autophagy is also part of cell death pathways in certain situations (11). Autophagy can be a player in apoptosis-independent type-2 cell death (type-1 cell death is apoptosis), also known as autophagic cell death. This situation has been shown to occur when the apoptotic machinery is crippled in mammalian cells (12, 13). Autophagy can also be part of the apoptotic program, for instance in tumor necrosis factor-α-induced cell death when NF-κB is inhibited (14), or in human immunodeficiency virus envelope-mediated cell death in bystander naive CD4 T cells (15). Moreover autophagy has recently been shown to be required for the externalization of phosphatidylserine, the eat-me signal for phagocytic cells, at the surface of apoptotic cells (16).The complex relationship between autophagy and apoptosis reflects the intertwined regulation of these processes (17, 18). Many signaling pathways involved in the regulation of autophagy also regulate apoptosis. This intertwining has recently been shown to occur at the level of the molecular machinery of autophagy. In fact the anti-apoptotic protein Bcl-2 has been shown to inhibit starvation-induced autophagy by interacting with the autophagy protein Beclin 1 (19). Beclin 1 is one of the Atg proteins conserved from yeast to humans (it is the mammalian orthologue of yeast Atg6) and is involved in autophagosome formation (20). Beclin 1 is a platform protein that interacts with several different partners, including hVps34 (class III phosphatidylinositol 3-kinase), which is responsible for the synthesis of phosphatidylinositol 3-phosphate. The production of this lipid is important for events associated with the nucleation of the isolation membrane before it elongates and closes to form autophagosomes in response to other Atg proteins, including the Atg12 and LC32 (microtubule-associated protein light chain 3 is the mammalian orthologue of the yeast Atg8) ubiquitin-like conjugation systems (3, 21). Various partners associated with the Beclin 1 complex modulate the activity of hVps34. For instance, Bcl-2 inhibits the activity of this enzyme, whereas UVRAG, Ambra-1, and Bif-1 all up-regulate it (22, 23).In view of the intertwining between autophagy and apoptosis, it is noteworthy that Beclin 1 belongs to the BH3-only family of proteins (2426). However, and unlike most of the proteins in this family, Beclin 1 is not able to trigger apoptosis when its expression is forced in cells (27). A BH3-mimetic drug, ABT-737, is able to dissociate the Beclin 1-Bcl-2 complex, and to trigger autophagy by mirroring the effect of starvation (25).The sphingolipids constitute a family of bioactive lipids (2832) of which several members, such as ceramide and sphingosine 1-phosphate, are signaling molecules. These molecules constitute a “sphingolipid rheostat” that determines the fate of the cell, because in many settings ceramide is pro-apoptotic and sphingosine 1-phosphate mitigates this apoptotic effect (31, 32). However, ceramide is also engaged in a wide variety of other cell processes, such as the formation of exosomes (33), differentiation, cell proliferation, and senescence (34). Recently we showed that both ceramide and sphingosine 1-phosphate are able to stimulate autophagy (35, 36). It has also been shown that ceramide triggers autophagy in a large panel of mammalian cells (3739). However, elucidation of the mechanism by which ceramide stimulates autophagy is still in its infancy. We have previously demonstrated that ceramide induces autophagy in breast and colon cancer cells by inhibiting the Class I phosphatidylinositol 3-phosphate/mTOR signaling pathway, which plays a central role in inhibiting autophagy (36). Inhibition of mTOR is another hallmark of starvation-induced autophagy (17). This finding led us to investigate the effect of ceramide on the Beclin 1-Bcl-2 complex. The results presented here show that ceramide is more potent than starvation in dissociating the Beclin 1-Bcl-2 complex (see Ref. 40). This dissociation is dependent on three phosphorylation sites (Thr69, Ser70, and Ser87) located in a non-structural loop of Bcl-2. Ceramide induces the c-Jun N-terminal kinase 1-dependent phosphorylation of Bcl-2. Expression of a dominant negative form of JNK1 blocks Bcl-2 phosphorylation, and thus the induction of autophagy by ceramide. These findings help to explain how autophagy is regulated by a major lipid second messenger.  相似文献   

15.
The activation of LFA-1 (lymphocyte function-associated antigen) is a critical event for T cell co-stimulation. The mechanism of LFA-1 activation involves both affinity and avidity regulation, but the role of each in T cell activation remains unclear. We have identified antibodies that recognize and block different affinity states of the mouse LFA-1 I-domain. Monoclonal antibody 2D7 preferentially binds to the low affinity conformation, and this specific binding is abolished when LFA-1 is locked in the high affinity conformation. In contrast, M17/4 can bind both the locked high and low affinity forms of LFA-1. Although both 2D7 and M17/4 are blocking antibodies, 2D7 is significantly less potent than M17/4 in blocking LFA-1-mediated adhesion; thus, blocking high affinity LFA-1 is critical for preventing LFA-1-mediated adhesion. Using these reagents, we investigated whether LFA-1 affinity regulation affects T cell activation. We found that blocking high affinity LFA-1 prevents interleukin-2 production and T cell proliferation, demonstrated by TCR cross-linking and antigen-specific stimulation. Furthermore, there is a differential requirement of high affinity LFA-1 in the activation of CD4+ and CD8+ T cells. Although CD4+ T cell activation depends on both high and low affinity LFA-1, only high affinity LFA-1 provides co-stimulation for CD8+ T cell activation. Together, our data demonstrated that the I-domain of LFA-1 changes to the high affinity state in primary T cells, and high affinity LFA-1 is critical for facilitating T cell activation. This implicates LFA-1 activation as a novel regulatory mechanism for the modulation of T cell activation and proliferation.LFA-1 (lymphocyte function-associated antigen), an integrin family member, is important in regulating leukocyte adhesion and T cell activation (1, 2). LFA-1 consists of the αL (CD11a) and β2 (CD18) heterodimer. The ligands for LFA-1, including intercellular adhesion molecule ICAM3-1, ICAM-2, and ICAM-3, are expressed on antigen-presenting cells (APCs), endothelial cells, and lymphocytes (1). Mice that are deficient in LFA-1 have defects in leukocyte adhesion, lymphocyte proliferation, and tumor rejection (35). Blocking LFA-1 with antibodies can prevent inflammation, autoimmunity, organ graft rejection, and graft versus host disease in human and murine models (610).LFA-1 is constitutively expressed on the surface of leukocytes in an inactive state. Activation of LFA-1 is mediated by inside-out signals from the cytoplasm (1, 11). Subsequently, activated LFA-1 binds to the ligands and transduces outside-in signals back into the cytoplasm that result in cell adhesion and activation (12, 13). The activation of LFA-1 is a critical event in the formation of the immunological synapse, which is important for T cell activation (2, 14, 15). The active state of LFA-1 is regulated by chemokines and the T cell receptor (TCR) through Rap1 signaling (16). LFA-1 ligation lowers the activation threshold and affects polarization in CD4+ T cells (17). Moreover, productive LFA-1 engagement facilitates efficient activation of cytotoxic T lymphocytes and initiates a distinct signal essential for the effector function (1820). Thus, LFA-1 activation is essential for the optimal activation of T cells.The mechanism of LFA-1 activation involves both affinity (conformational changes within the molecule) and avidity (receptor clustering) regulation (2123). The I-domain of the LFA-1 αL subunit is the primary ligand-binding site and has been proposed to change conformation, leading to an increased affinity for ligands (2426). The structural basis of the conformational changes in the I-domain of LFA-1 has been extensively characterized (27). Previously, we have demonstrated that the conformation of the LFA-1 I-domain changes from the low affinity to the high affinity state upon activation. By introducing disulfide bonds into the I-domain, LFA-1 can be locked in either the closed or open conformation, which represents the “low affinity” or “high affinity” state, respectively (28, 29). In addition, we identified antibodies that are sensitive to the affinity changes in the I-domain of human LFA-1 and showed that the activation-dependent epitopes are exposed upon activation (30). This study supports the presence of the high affinity conformation upon LFA-1 activation in cell lines. It has been demonstrated recently that therapeutic antagonists, such as statins, inhibit LFA-1 activation and immune responses by locking LFA-1 in the low affinity state (3134). Furthermore, high affinity LFA-1 has been shown to be important for mediating the adhesion of human T cells (35, 36). Thus, the affinity regulation is a critical step in LFA-1 activation.LFA-1 is a molecule of great importance in the immune system, and its activation state influences the outcome of T cell activation. Our previous data using the activating LFA-1 I-domain-specific antibody MEM83 indicate that avidity and affinity of the integrin can be coupled during activation (37). However, whether affinity or avidity regulation of LFA-1 contributes to T cell activation remains controversial (23, 38, 39). Despite the recent progress suggesting that conformational changes represent a key step in the activation of LFA-1, there are considerable gaps to be filled. When LFA-1 is activated, the subsequent outside-in signaling contributes to T cell activation via immunological synapse and LFA-1-dependent signaling. It is critical to determine whether high affinity LFA-1 participates in the outside-in signaling and affects the cellular activation of T cells. Nevertheless, the rapid and dynamic process of LFA-1 activation has hampered further understanding of the role of high affinity LFA-1 in primary T cell activation. The affinity of LFA-1 for ICAM-1 increases up to 10,000-fold within seconds and involves multiple reversible steps (23). In addition, the activation of LFA-1 regulates both adhesion and activation of T cells, two separate yet closely associated cellular functions. When LFA-1 is constitutively expressed in the active state in mice, immune responses are broadly impaired rather than hyperactivated, suggesting the complexity of affinity regulation (40). Therefore, it is difficult to dissect the mechanisms by which high affinity LFA-1 regulates stepwise activation of T cells in the whole animal system.In the present study, we identified antibodies recognizing and blocking different affinity states of mouse LFA-1. These reagents allowed us to determine the role of affinity regulation in T cell activation. We found that blocking high affinity LFA-1 inhibited IL-2 production and proliferation in T cells. Furthermore, there is a differential requirement of high affinity LFA-1 in antigen-specific activation of CD4+ and CD8+ T cells. The activation of CD4+ T cells depends on both high and low affinity LFA-1. For CD8+ T cell activation, only high affinity LFA-1 provides co-stimulation. Thus, affinity regulation of LFA-1 is critical for the activation and proliferation of naive T cells.  相似文献   

16.
17.
Rapid protein kinase D (PKD) activation and phosphorylation via protein kinase C (PKC) have been extensively documented in many cell types cells stimulated by multiple stimuli. In contrast, little is known about the role and mechanism(s) of a recently identified sustained phase of PKD activation in response to G protein-coupled receptor agonists. To elucidate the role of biphasic PKD activation, we used Swiss 3T3 cells because PKD expression in these cells potently enhanced duration of ERK activation and DNA synthesis in response to Gq-coupled receptor agonists. Cell treatment with the preferential PKC inhibitors GF109203X or Gö6983 profoundly inhibited PKD activation induced by bombesin stimulation for <15 min but did not prevent PKD catalytic activation induced by bombesin stimulation for longer times (>60 min). The existence of sequential PKC-dependent and PKC-independent PKD activation was demonstrated in 3T3 cells stimulated with various concentrations of bombesin (0.3–10 nm) or with vasopressin, a different Gq-coupled receptor agonist. To gain insight into the mechanisms involved, we determined the phosphorylation state of the activation loop residues Ser744 and Ser748. Transphosphorylation targeted Ser744, whereas autophosphorylation was the predominant mechanism for Ser748 in cells stimulated with Gq-coupled receptor agonists. We next determined which phase of PKD activation is responsible for promoting enhanced ERK activation and DNA synthesis in response to Gq-coupled receptor agonists. We show, for the first time, that the PKC-independent phase of PKD activation mediates prolonged ERK signaling and progression to DNA synthesis in response to bombesin or vasopressin through a pathway that requires epidermal growth factor receptor-tyrosine kinase activity. Thus, our results identify a novel mechanism of Gq-coupled receptor-induced mitogenesis mediated by sustained PKD activation through a PKC-independent pathway.The understanding of the mechanisms that control cell proliferation requires the identification of the molecular pathways that govern the transition of quiescent cells into the S phase of the cell cycle. In this context the activation and phosphorylation of protein kinase D (PKD),4 the founding member of a new protein kinase family within the Ca2+/calmodulin-dependent protein kinase (CAMK) group and separate from the previously identified PKCs (for review, see Ref. 1), are attracting intense attention. In unstimulated cells, PKD is in a state of low catalytic (kinase) activity maintained by autoinhibition mediated by the N-terminal domain, a region containing a repeat of cysteinerich zinc finger-like motifs and a pleckstrin homology (PH) domain (14). Physiological activation of PKD within cells occurs via a phosphorylation-dependent mechanism first identified in our laboratory (57). In response to cellular stimuli (1), including phorbol esters, growth factors (e.g. PDGF), and G protein-coupled receptor (GPCR) agonists (6, 816) that signal through Gq, G12, Gi, and Rho (11, 1519), PKD is converted into a form with high catalytic activity, as shown by in vitro kinase assays performed in the absence of lipid co-activators (5, 20).During these studies multiple lines of evidence indicated that PKC activity is necessary for rapid PKD activation within intact cells. For example, rapid PKD activation was selectively and potently blocked by cell treatment with preferential PKC inhibitors (e.g. GF109203X or Gö6983) that do not directly inhibit PKD catalytic activity (5, 20), implying that PKD activation in intact cells is mediated directly or indirectly through PKCs. Many reports demonstrated the operation of a rapid PKC/PKD signaling cascade induced by multiple GPCR agonists and other receptor ligands in a range of cell types (for review, see Ref. 1). Our previous studies identified Ser744 and Ser748 in the PKD activation loop (also referred as activation segment or T-loop) as phosphorylation sites critical for PKC-mediated PKD activation (1, 4, 7, 17, 21). Collectively, these findings demonstrated the existence of a rapidly activated PKC-PKD protein kinase cascade(s). In a recent study we found that the rapid PKC-dependent PKD activation was followed by a late, PKC-independent phase of catalytic activation and phosphorylation induced by stimulation of the bombesin Gq-coupled receptor ectopically expressed in COS-7 cells (22). This study raised the possibility that PKD mediates rapid biological responses downstream of PKCs, whereas, in striking contrast, PKD could mediate long term responses through PKC-independent pathways. Despite its potential importance for defining the role of PKC and PKD in signal transduction, this hypothesis has not been tested in any cell type.Accumulating evidence demonstrates that PKD plays an important role in several cellular processes and activities, including signal transduction (14, 2325), chromatin organization (26), Golgi function (27, 28), gene expression (2931), immune regulation (26), and cell survival, adhesion, motility, differentiation, DNA synthesis, and proliferation (for review, see Ref. 1). In Swiss 3T3 fibroblasts, a cell line used extensively as a model system to elucidate mechanisms of mitogenic signaling (3234), PKD expression potently enhances ERK activation, DNA synthesis, and cell proliferation induced by Gq-coupled receptor agonists (8, 14). Here, we used this model system to elucidate the role and mechanism(s) of biphasic PKD activation. First, we show that the Gq-coupled receptor agonists bombesin and vasopressin, in contrast to phorbol esters, specifically induce PKD activation through early PKC-dependent and late PKC-independent mechanisms in Swiss 3T3 cells. Subsequently, we demonstrate for the first time that the PKC-independent phase of PKD activation is responsible for promoting ERK signaling and progression to DNA synthesis through an epidermal growth factor receptor (EGFR)-dependent pathway. Thus, our results identify a novel mechanism of Gq-coupled receptor-induced mitogenesis mediated by sustained PKD activation through a PKC-independent pathway.  相似文献   

18.
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   

19.
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   

20.
ATP-binding cassette (ABC) transporters transduce the free energy of ATP hydrolysis to power the mechanical work of substrate translocation across cell membranes. MsbA is an ABC transporter implicated in trafficking lipid A across the inner membrane of Escherichia coli. It has sequence similarity and overlapping substrate specificity with multidrug ABC transporters that export cytotoxic molecules in humans and prokaryotes. Despite rapid advances in structure determination of ABC efflux transporters, little is known regarding the location of substrate-binding sites in the transmembrane segment and the translocation pathway across the membrane. In this study, we have mapped residues proximal to the daunorubicin (DNR)-binding site in MsbA using site-specific, ATP-dependent quenching of DNR intrinsic fluorescence by spin labels. In the nucleotide-free MsbA intermediate, DNR-binding residues cluster at the cytoplasmic end of helices 3 and 6 at a site accessible from the membrane/water interface and extending into an aqueous chamber formed at the interface between the two transmembrane domains. Binding of a nonhydrolyzable ATP analog inverts the transporter to an outward-facing conformation and relieves DNR quenching by spin labels suggesting DNR exclusion from proximity to the spin labels. The simplest model consistent with our data has DNR entering near an elbow helix parallel to the water/membrane interface, partitioning into the open chamber, and then translocating toward the periplasm upon ATP binding.ATP-binding cassette (ABC)2 transporters transduce the energy of ATP hydrolysis to power the movement of a wide range of substrates across the cell membranes (1, 2). They constitute the largest family of prokaryotic transporters, import essential cell nutrients, flip lipids, and export toxic molecules (3). Forty eight human ABC transporters have been identified, including ABCB1, or P-glycoprotein, which is implicated in cross-resistance to drugs and cytotoxic molecules (4, 5). Inherited mutations in these proteins are linked to diseases such as cystic fibrosis, persistent hypoglycemia of infancy, and immune deficiency (6).The functional unit of an ABC transporter consists of four modules. Two highly conserved ABCs or nucleotide-binding domains (NBDs) bind and hydrolyze ATP to supply the active energy for transport (7). ABCs drive the mechanical work of proteins with diverse functions ranging from membrane transport to DNA repair (3, 5). Substrate specificity is determined by two transmembrane domains (TMDs) that also provide the translocation pathway across the bilayer (7). Bacterial ABC exporters are expressed as monomers, each consisting of one NBD and one TMD, that dimerize to form the active transporter (3). The number of transmembrane helices and their organization differ significantly between ABC importers and exporters reflecting the divergent structural and chemical nature of their substrates (1, 8, 9). Furthermore, ABC exporters bind substrates directly from the cytoplasm or bilayer inner leaflet and release them to the periplasm or bilayer outer leaflet (10, 11). In contrast, bacterial importers have their substrates delivered to the TMD by a dedicated high affinity substrate-binding protein (12).In Gram-negative bacteria, lipid A trafficking from its synthesis site on the inner membrane to its final destination in the outer membrane requires the ABC transporter MsbA (13). Although MsbA has not been directly shown to transport lipid A, suppression of MsbA activity leads to cytoplasmic accumulation of lipid A and inhibits bacterial growth strongly suggesting a role in translocation (14-16). In addition to this role in lipid A transport, MsbA shares sequence similarity with multidrug ABC transporters such as human ABCB1, LmrA of Lactococcus lactis, and Sav1866 of Staphylococcus aureus (16-19). ABCB1, a prototype of the ABC family, is a plasma membrane protein whose overexpression provides resistance to chemotherapeutic agents in cancer cells (1). LmrA and MsbA have overlapping substrate specificity with ABCB1 suggesting that both proteins can function as drug exporters (18, 20). Indeed, cells expressing MsbA confer resistance to erythromycin and ethidium bromide (21). MsbA can be photolabeled with the ABCB1/LmrA substrate azidopine and can transport Hoechst 33342 (H33342) across membrane vesicles in an energy-dependent manner (21).The structural mechanics of ABC exporters was revealed from comparison of the MsbA crystal structures in the apo- and nucleotide-bound states as well as from analysis by spin labeling EPR spectroscopy in liposomes (17, 19, 22, 23). The energy harnessed from ATP binding and hydrolysis drives a cycle of NBD association and dissociation that is transmitted to induce reorientation of the TMD from an inward- to outward-facing conformation (17, 19, 22). Large amplitude motion closes the cytoplasmic end of a chamber found at the interface between the two TMDs and opens it to the periplasm (23). These rearrangements lead to significant changes in chamber hydration, which may drive substrate translocation (22).Substrate binding must precede energy input, otherwise the cycle is futile, wasting the energy of ATP hydrolysis without substrate extrusion (7). Consistent with this model, ATP binding reduces ABCB1 substrate affinity, potentially through binding site occlusion (24-26). Furthermore, the TMD substrate-binding event signals the NBD to stimulate ATP hydrolysis increasing transport efficiency (1, 27, 28). However, there is a paucity of information regarding the location of substrate binding, the transport pathway, and the structural basis of substrate recognition by ABC exporters. In vitro studies of MsbA substrate specificity identify a broad range of substrates that stimulate ATPase activity (29). In addition to the putative physiological substrates lipid A and lipopolysaccharide (LPS), the ABCB1 substrates Ilmofosine, H33342, and verapamil differentially enhance ATP hydrolysis of MsbA (29, 30). Intrinsic MsbA tryptophan (Trp) fluorescence quenching by these putative substrate molecules provides further support of interaction (29).Extensive biochemical analysis of ABCB1 and LmrA provides a general model of substrate binding to ABC efflux exporters. This so-called “hydrophobic cleaner model” describes substrates binding from the inner leaflet of the bilayer and then translocating through the TMD (10, 31, 32). These studies also identified a large number of residues involved in substrate binding and selectivity (33). When these crucial residues are mapped onto the crystal structures of MsbA, a subset of homologous residues clusters to helices 3 and 6 lining the putative substrate pathway (34). Consistent with a role in substrate binding and specificity, simultaneous replacement of two serines (Ser-289 and Ser-290) in helix 6 of MsbA reduces binding and transport of ethidium and taxol, although H33342 and erythromycin interactions remain unaffected (34).The tendency of lipophilic substrates to partition into membranes confounds direct analysis of substrate interactions with ABC exporters (35, 36). Such partitioning may promote dynamic collisions with exposed Trp residues and nonspecific cross-linking in photo-affinity labeling experiments. In this study, we utilize a site-specific quenching approach to identify residues in the vicinity of the daunorubicin (DNR)-binding site (37). Although the data on DNR stimulation of ATP hydrolysis is inconclusive (20, 29, 30), the quenching of MsbA Trp fluorescence suggests a specific interaction. Spin labels were introduced along transmembrane helices 3, 4, and 6 of MsbA to assess their ATP-dependent quenching of DNR fluorescence. Residues that quench DNR cluster along the cytoplasmic end of helices 3 and 6 consistent with specific binding of DNR. Furthermore, many of these residues are not lipid-exposed but face the putative substrate chamber formed between the two TMDs. These residues are proximal to two Trps, which likely explains the previously reported quenching (29). Our results suggest DNR partitions to the membrane and then binds MsbA in a manner consistent with the hydrophobic cleaner model. Interpretation in the context of the crystal structures of MsbA identifies a putative translocation pathway through the transmembrane segment.  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