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.
Membrane fusion without lysis has been reconstituted with purified yeast vacuolar SNAREs (soluble N-ethylmaleimide-sensitive factor attachment protein receptors), the SNARE chaperones Sec17p/Sec18p and the multifunctional HOPS complex, which includes a subunit of the SNARE-interactive Sec1-Munc18 family, and vacuolar lipids: phosphatidylcholine (PC), phosphatidylethanolamine (PE), phosphatidylinositol (PI), phosphatidylserine (PS), phosphatidic acid (PA), cardiolipin (CL), ergosterol (ERG), diacylglycerol (DAG), and phosphatidylinositol 3-phosphate (PI3P). We now report that many of these lipids are required for rapid and efficient fusion of the reconstituted SNARE proteoliposomes in the presence of SNARE chaperones. Omission of either PE, PA, or PI3P from the complete set of lipids strongly reduces fusion, and PC, PE, PA, and PI3P constitute a minimal set of lipids for fusion. PA could neither be replaced by other lipids with small headgroups such as DAG or ERG nor by the acidic lipids PS or PI. PA is needed for full association of HOPS and Sec18p with proteoliposomes having a minimal set of lipids. Strikingly, PA and PE are as essential for SNARE complex assembly as for fusion, suggesting that these lipids facilitate functional interactions among SNAREs and SNARE chaperones.Biological membrane fusion is the regulated rearrangement of the lipids in two apposed sealed membranes to form one bilayer while mixing lumenal contents without leakage or lysis. It is fundamental for intracellular vesicular traffic, cell growth and division, regulated secretion of hormones and other blood proteins, and neurotransmission and thus has attracted wide and sustained study (1, 2). Its fundamental mechanisms are conserved and employ a Rab-family GTPase, proteins which bind to the GTP-bound form of a Rab, termed its “effectors” (3), and SNARE3 (soluble N-ethylmaleimide-sensitive factor attachment protein receptors) proteins (4) with their attendant chaperones. SNAREs are integral or peripheral membrane proteins with characteristic heptad-repeat domains, which can associate in 4-helical coiled-coils (5), termed “cis-SNARE complexes,” if they are all anchored to the same membrane bilayer, or “trans-SNARE complexes” if they are anchored to apposed membranes.Stable membrane proximity (docking) does not suffice for fusion. Studies in model systems have shown that fusion can be promoted by any of several agents, which promote bilayer rearrangement, such as diacylglycerol (6), high levels of calcium (7), viral-encoded fusion proteins (8, 9), or SNAREs (10, 11). These studies frequently employed liposomes or proteoliposomes of simple lipid composition, suggesting that fusion may not have stringent requirements of lipid head group species. However, each of these model fusion reactions is accompanied by substantial lysis (1215), whereas the preservation of subcellular compartments is a hallmark of physiological membrane fusion.We have studied membrane fusion with the vacuole (lysosome) of Saccharomyces cerevisiae (reviewed in Ref. 16). The fusion of isolated vacuoles requires the Rab Ypt7p, 4 SNAREs (Vam3p, Vti1p, Vam7p, and Nyv1p), the SNARE chaperones Sec17p (α-soluble N-ethylmaleimide-sensitive factor attachment protein)/Sec18p (N-ethylmaleimide-sensitive factor) and the hexameric HOPS complex (17), and key “regulatory” lipids including ERG, phosphoinositides, and DAG (18). HOPS interacts physically or functionally with each component of this fusion system. HOPS stably associates with Ypt7p in its GTP-bound state (19). One HOPS subunit, Vps33p, is a member of the Sec1-Munc18 family of SNARE-binding proteins, and HOPS exhibits direct affinity for SNAREs (17, 2022) and proofreads correct vacuolar SNARE pairing (23). HOPS also has direct affinity for phosphoinositides (17). The SNAREs on isolated vacuoles are in cis-complexes, which are disassembled by Sec17p, Sec18p, and ATP (24). Docking requires Ypt7p (25) and HOPS (17). During docking, vacuoles are drawn against each other until each has a substantial membrane domain tightly apposed to the other. Each of the proteins (26) and lipids (18) required for fusion becomes enriched in a ring-shaped microdomain, the “vertex ring,” which surrounds the two tightly apposed membrane domains. Not only do the proteins depend on each other, in a cascade fashion, for vertex ring enrichment, and the lipids depend on each other for their vertex ring enrichment as well, but the lipids and proteins are mutually interdependent for their enrichment at this ring-shaped microdomain (18, 27). Fusion occurs around the ring, joining the two organelles. The fusion of vacuoles bearing physiological fusion constituents does not cause measurable organelle lysis, although fusion supported exclusively by higher levels of SNARE proteins is accompanied by massive lysis (28), in accord with model liposome studies (14). Thus fusion microdomain assembly and the coordinate action of SNAREs with other proteins and lipids to promote fusion without lysis are central topics in membrane fusion studies.Reconstitution of fusion with pure components allows chemical definition of essential elements of this biologically important reaction. Although SNAREs can drive a slow fusion of PC/PS proteoliposomes (29), this was not stimulated by HOPS and Sec17p/Sec18p (30). SNARE proteoliposomes bearing all the vacuolar lipids (18, 3133), PC, PE, PI, PS, CL, PA, ERG, DAG, PI3P, and phosphatidylinositol 4,5-bisphosphate (PI(4,5)P2), showed rapid and efficient fusion that was fully dependent on Sec17p/Sec18p and HOPS (30). The omission of either DAG, ERG, or phosphoinositide from the liposomes caused a marked reduction in fusion (30). We now report that PE and PA are also necessary for rapid and efficient fusion, function in distinct manners, and are required for efficient assembly of newly formed SNARE complexes by the SNARE chaperones Sec17p/Sec18p and HOPS.  相似文献   

2.
Insulin plays a central role in the regulation of vertebrate metabolism. The hormone, the post-translational product of a single-chain precursor, is a globular protein containing two chains, A (21 residues) and B (30 residues). Recent advances in human genetics have identified dominant mutations in the insulin gene causing permanent neonatal-onset DM2 (14). The mutations are predicted to block folding of the precursor in the ER of pancreatic β-cells. Although expression of the wild-type allele would in other circumstances be sufficient to maintain homeostasis, studies of a corresponding mouse model (57) suggest that the misfolded variant perturbs wild-type biosynthesis (8, 9). Impaired β-cell secretion is associated with ER stress, distorted organelle architecture, and cell death (10). These findings have renewed interest in insulin biosynthesis (1113) and the structural basis of disulfide pairing (1419). Protein evolution is constrained not only by structure and function but also by susceptibility to toxic misfolding.Insulin plays a central role in the regulation of vertebrate metabolism. The hormone, the post-translational product of a single-chain precursor, is a globular protein containing two chains, A (21 residues) and B (30 residues). Recent advances in human genetics have identified dominant mutations in the insulin gene causing permanent neonatal-onset DM2 (14). The mutations are predicted to block folding of the precursor in the ER of pancreatic β-cells. Although expression of the wild-type allele would in other circumstances be sufficient to maintain homeostasis, studies of a corresponding mouse model (57) suggest that the misfolded variant perturbs wild-type biosynthesis (8, 9). Impaired β-cell secretion is associated with ER stress, distorted organelle architecture, and cell death (10). These findings have renewed interest in insulin biosynthesis (1113) and the structural basis of disulfide pairing (1419). Protein evolution is constrained not only by structure and function but also by susceptibility to toxic misfolding.  相似文献   

3.
The visual photoreceptor rhodopsin is a prototypical class I (rhodopsin-like) G protein-coupled receptor. Photoisomerization of the covalently bound ligand 11-cis-retinal leads to restructuring of the cytosolic face of rhodopsin. The ensuing protonation of Glu-134 in the class-conserved D(E)RY motif at the C-terminal end of transmembrane helix-3 promotes the formation of the G protein-activating state. Using transmembrane segments derived from helix-3 of bovine rhodopsin, we show that lipid protein interactions play a key role in this cytosolic “proton switch.” Infrared and fluorescence spectroscopic pKa determinations reveal that the D(E)RY motif is an autonomous functional module coupling side chain neutralization to conformation and helix positioning as evidenced by side chain to lipid headgroup Foerster resonance energy transfer. The free enthalpies of helix stabilization and hydrophobic burial of the neutral carboxyl shift the side chain pKa into the range typical of Glu-134 in photoactivated rhodopsin. The lipid-mediated coupling mechanism is independent of interhelical contacts allowing its conservation without interference with the diversity of ligand-specific interactions in class I G protein-coupled receptors.G protein-coupled receptors (GPCRs)2 are hepta-helical membrane proteins that couple a large variety of extracellular signals to cell-specific responses via activation of G proteins. In the visual photoreceptor rhodopsin, a prototypical class I GPCR (1, 2), molecular activation processes can be monitored in real time by spectroscopic assays and analyzed in the context of several crystal structures (38). The primary signal for rhodopsin is the 11-cis to all-trans photoisomerization of retinal covalently bound to the apoprotein opsin through a protonated Schiff base to Lys296. Current models converge toward a picture in which “microdomains” act as conformational switches that are coupled to different degrees to the primary activation process. Two activating “proton switches” have been identified (9) as follows: breakage of an intramolecular salt bridge (10) by transfer of the Schiff base proton to its counter ion Glu-113 (11) is followed by movement of helix-6 (H6) (12, 13) in the metarhodopsin IIa (MIIa) to MIIb transition. The MIIb state takes up a proton at Glu-134 (14) in the class-conserved D(E)RY motif at the C-terminal end of helix-3 (H3) leading to the MIIbH+ intermediate (15, 16), which activates transducin (Gt), the G protein of the photoreceptor cell. Glu-134 regulates the pH sensitivity of receptor signaling (17) in membranes as reviewed previously (18), and in complex with Gt the protonated state of the carboxyl group becomes stabilized (19). This charge alteration is linked to the release of an “ionic lock,” originally described for the β2-adrenergic receptor (20), which also in rhodopsin stabilizes the inactive state (16) through interactions between the cytosolic ends of H3 and H6 (21).In the absence of a lipidic bilayer, proton uptake and H6 movement become uncoupled (15). Lipidic composition affects MII formation, rhodopsin structure, and oligomerization (2224) and differs at the rhodopsin membrane interface from the bulk lipidic phase (25). Likewise, MII formation specifically affects lipid structure (26). Although of fundamental importance for GPCR activation, the potential implication of lipid protein interactions in “proton switching” is not clear. A functional role of Glu-134 in lipid interactions has been originally derived from IR spectra where E134Q replacement abolished changes of lipid headgroup vibrations in the MIIGt complex (19). Computational approaches emphasized the “strategic” location of the D(E)RY motif (27), and the Glu-134 carboxyl pKa may critically depend on the lipid protein interface (28). However, the implications for proton switching are not evident, and the theoretical interest is contrasted by the lack of experimental data addressing the effect of the lipidic phase on side chain protonation, secondary structure, and membrane topology of the D(E)RY motif.We have studied the coupling between conformation and protonation in single transmembrane segments derived from H3 of bovine rhodopsin. We have assessed the “modular” function of the D(E)RY motif by determining parameters not evident from the crystal structures, i.e. the pKa of the conserved carboxyl, its linkage to helical structure, and the effect of protonation on side chain to lipid headgroup distance. We show that the D(E)RY motif encodes an autonomous “proton switch” controlling side chain exposure and helix formation in the low dielectric of a lipidic phase. The data ascribe a functional role to lipid protein interactions that couple the chemical potential of protons to an activity-promoting GPCR conformation in a ligand-independent manner.  相似文献   

4.
Human polymerase kappa (hPol κ) is one of four eukaryotic Y-class DNA polymerases and may be an important element in the cellular response to polycyclic aromatic hydrocarbons such as benzo[a]pyrene, which can lead to reactive oxygenated metabolite-mediated oxidative stress. Here, we present a detailed analysis of the activity and specificity of hPol κ bypass opposite the major oxidative adduct 7,8-dihydro-8-oxo-2′-deoxyguanosine (8-oxoG). Unlike its archaeal homolog Dpo4, hPol κ bypasses this lesion in an error-prone fashion by inserting mainly dATP. Analysis of transient-state kinetics shows diminished “bursts” for dATP:8-oxoG and dCTP:8-oxoG incorporation, indicative of non-productive complex formation, but dATP:8-oxoG insertion events that do occur are 2-fold more efficient than dCTP:G insertion events. Crystal structures of ternary hPol κ complexes with adducted template-primer DNA reveal non-productive (dGTP and dATP) alignments of incoming nucleotide and 8-oxoG. Structural limitations placed upon the hPol κ by interactions between the N-clasp and finger domains combined with stabilization of the syn-oriented template 8-oxoG through the side chain of Met-135 both appear to contribute to error-prone bypass. Mutating Leu-508 in the little finger domain of hPol κ to lysine modulates the insertion opposite 8-oxoG toward more accurate bypass, similar to previous findings with Dpo4. Our structural and activity data provide insight into important mechanistic aspects of error-prone bypass of 8-oxoG by hPol κ compared with accurate and efficient bypass of the lesion by Dpo4 and polymerase η.DNA damage incurred by a multitude of endogenous and exogenous factors constitutes an inevitable challenge for the replication machinery, and various mechanisms exist to either remove the resulting lesions or bypass them in a more or less mutation-prone fashion (1). Error-prone polymerases are central to trans-lesion synthesis across sites of damaged DNA (2, 3). Four so-called Y-class DNA polymerases have been identified in humans, Pol η,4 Pol ι, Pol κ, and Rev1, which exhibit different activities and abilities to replicate past a flurry of individual lesions (4, 5). Homologs have also been identified and characterized in other organisms, notably DinB (Pol IV) in Escherichia coli (68), Dbh in Sulfolobus acidocaldarius (9, 10), and Dpo4 in Sulfolobus solfataricus (11, 12). A decade of investigations directed at the structural and functional properties of bypass polymerases have significantly improved our understanding of this class of enzymes (5, 13). A unique feature of Y-class polymerases, compared with the common right-handed arrangement of palm, thumb, and finger subdomains of high fidelity (i.e. A-class) DNA polymerases (14), is a “little finger” or “PAD” (palm-associated domain) subdomain that plays a crucial role in lesion bypass (12, 1521). In addition to the little finger subdomain at the C-terminal end of the catalytic core, both Rev1 and Pol κ exhibit an N-terminal extension that is absent in other translesion polymerases. The N-terminal extension in the structure of the ternary (human) hPol κ·DNA·dTTP complex folds into a U-shaped tether-helix-turn-helix “clasp” that is located between the thumb and little finger domains, allowing the polymerase to completely encircle the DNA (18). Although the precise role of the clasp for lesion bypass by hPol κ remains to be established, it is clear that this entity is functionally important, because mutant enzymes with partially or completely removed clasps exhibit diminished catalytic activity compared with the full-length catalytic core (hPol κ N1–526) or a core lacking the N-terminal 19 residues (hPol κ N19–526; the construct used for crystal structure determination of the ternary complex (18)).7,8-Dihydro-8-oxo-2′-deoxyguanosine (8-oxoG), found in both lower organisms and eukaryotes, is a major lesion that is a consequence of oxidative stress. The lesion is of relevance not only because of its association with cancer (22, 23), but also in connection with aging (24), hepatitis (25), and infertility (26). It is far from clear which DNA polymerases bypass 8-oxoG most often in a cellular context, but given the ubiquitous nature of the lesion it seems likely that more than one enzyme could encounter the lesion. Replicative polymerases commonly insert dATP opposite template 8-oxoG, with the lesion adopting the preferred syn conformation (e.g. 27, 28). It was recently found that the translesion polymerase Dpo4 from S. solfataricus synthesizes efficiently past 8-oxoG, inserting ≥95% dCTP > dATP opposite the lesion (29, 30). The efficient and low error bypass of the 8-oxoG lesion by Dpo4 is associated to a large extent with an arginine residue in the little finger domain (17). In the crystal structure of the ternary Dpo4·DNA·dCTP complex, the side chain of Arg-332 forms a hydrogen bond to the 8-oxygen of 8-oxoG, thus shifting the nucleoside conformational equilibrium toward the anti state and enabling a Watson-Crick binding mode with the incoming dCTP (30). The efficient and accurate replication of templates bearing 8-oxoG by yeast Pol η (31, 32) may indicate similarities between the bypass reactions catalyzed by the archaeal and eukaryotic enzymes. In contrast, bypass synthesis opposite 8-oxoG by human Pol κ is error-prone, resulting in efficient incorporation of A (3335). The inaccurate bypass of 8-oxoG is thought to contribute to the deleterious effects associated with the lesion. These observations indicate different behaviors of the eukaryotic trans-lesion Pol κ and its archaeal “homolog” Dpo4 vis-à-vis the major oxidative stress lesion 8-oxoG. A mechanistic understanding of human DNA polymerases that bypass 8-oxoG in an error-prone fashion, such as hPol κ, is therefore of great interest.To elucidate commonalities and differences between the trans-8-oxoG syntheses of S. solfataricus Dpo4, yeast Pol η, and hPol κ, we carried out a comprehensive analysis of the bypass activity for the latter with template·DNA containing the 8-oxoG lesion, including pre-steady-state and steady-state kinetics of primer extension opposite and beyond 8-oxoG and LC-MS/MS assays of full-length extension products. We determined crystal structures of ternary hPol κ-(8-oxoG)DNA-dGTP and hPol κ-(8-oxoG)DNA-dATP complexes, apparently the first for any complex with adducted DNA for the κ enzyme reported to date. Our work demonstrates clear distinctions between genetically related translesion polymerases and provides insights into the origins of the error-prone reactions opposite 8-oxoG catalyzed by Y-family DNA polymerases.  相似文献   

5.
Prion propagation involves a conformational transition of the cellular form of prion protein (PrPC) to a disease-specific isomer (PrPSc), shifting from a predominantly α-helical conformation to one dominated by β-sheet structure. This conformational transition is of critical importance in understanding the molecular basis for prion disease. Here, we elucidate the conformational properties of a disulfide-reduced fragment of human PrP spanning residues 91–231 under acidic conditions, using a combination of heteronuclear NMR, analytical ultracentrifugation, and circular dichroism. We find that this form of the protein, which similarly to PrPSc, is a potent inhibitor of the 26 S proteasome, assembles into soluble oligomers that have significant β-sheet content. The monomeric precursor to these oligomers exhibits many of the characteristics of a molten globule intermediate with some helical character in regions that form helices I and III in the PrPC conformation, whereas helix II exhibits little evidence for adopting a helical conformation, suggesting that this region is a likely source of interaction within the initial phases of the transformation to a β-rich conformation. This precursor state is almost as compact as the folded PrPC structure and, as it assembles, only residues 126–227 are immobilized within the oligomeric structure, leaving the remainder in a mobile, random-coil state.Prion diseases, such as Creutzfeldt-Jacob and Gerstmann-Sträussler-Scheinker in humans, scrapie in sheep, and bovine spongiform encephalopathy in cattle, are fatal neurological disorders associated with the deposition of an abnormally folded form of a host-encoded glycoprotein, prion (PrP)2 (1). These diseases may be inherited, arise sporadically, or be acquired through the transmission of an infectious agent (2, 3). The disease-associated form of the protein, termed the scrapie form or PrPSc, differs from the normal cellular form (PrPC) through a conformational change, resulting in a significant increase in the β-sheet content and protease resistance of the protein (3, 4). PrPC, in contrast, consists of a predominantly α-helical structured domain and an unstructured N-terminal domain, which is capable of binding a number of divalent metals (512). A single disulfide bond links two of the main α-helices and forms an integral part of the core of the structured domain (13, 14).According to the protein-only hypothesis (15), the infectious agent is composed of a conformational isomer of PrP (16) that is able to convert other isoforms to the infectious isomer in an autocatalytic manner. Despite numerous studies, little is known about the mechanism of conversion of PrPC to PrPSc. The most coherent and general model proposed thus far is that PrPC fluctuates between the dominant native state and minor conformations, one or a set of which can self-associate in an ordered manner to produce a stable supramolecular structure composed of misfolded PrP monomers (3, 17). This stable, oligomeric species can then bind to, and stabilize, rare non-native monomer conformations that are structurally complementary. In this manner, new monomeric chains are recruited and the system can propagate.In view of the above model, considerable effort has been devoted to generating and characterizing alternative, possibly PrPSc-like, conformations in the hope of identifying common properties or features that facilitate the formation of amyloid oligomers. This has been accomplished either through PrPSc-dependent conversion reactions (1820) or through conversion of PrPC in the absence of a PrPSc template (2125). The latter approach, using mainly disulfide-oxidized recombinant PrP, has generated a wide range of novel conformations formed under non-physiological conditions where the native state is relatively destabilized. These conformations have ranged from near-native (14, 26, 27), to those that display significant β-sheet content (21, 23, 2833). The majority of these latter species have shown a high propensity for aggregation, although not all are on-pathway to the formation of amyloid. Many of these non-native states also display some of the characteristics of PrPSc, such as increased β-sheet content, protease resistance, and a propensity for oligomerization (28, 29, 31) and some have been claimed to be associated with the disease process (34).One such PrP folding intermediate, termed β-PrP, differs from the majority of studied PrP intermediate states in that it is formed by refolding the PrP molecule from the native α-helical conformation (here termed α-PrP), at acidic pH in a reduced state, with the disulfide bond broken (22, 35). Although no covalent differences between the PrPC and PrPSc have been consistently identified to date, the role of the disulfide bond in prion propagation remains disputed (25, 3639). β-PrP is rich in β-sheet structure (22, 35), and displays many of the characteristics of a PrPSc-like precursor molecule, such as partial resistance to proteinase K digestion, and the ability to form amyloid fibrils in the presence of physiological concentrations of salts (40).The β-PrP species previously characterized, spanning residues 91–231 of PrP, was soluble at low ionic strength buffers and monomeric, according to elution volume on gel filtration (22). NMR analysis showed that it displayed radically different spectra to those of α-PrP, with considerably fewer observable peaks and markedly reduced chemical shift dispersion. Data from circular dichroism experiments showed that fixed side chain (tertiary) interactions were lost, in contrast to the well defined β-sheet secondary structure, and thus in conjunction with the NMR data, indicated that β-PrP possessed a number of characteristics associated with a “molten globule” folding intermediate (22). Such states have been proposed to be important in amyloid and fibril formation (41). Indeed, antibodies raised against β-PrP (e.g. ICSM33) are capable of recognizing native PrPSc (but not PrPC) (4244). Subsequently, a related study examining the role of the disulfide bond in PrP folding confirmed that a monomeric molten globule-like form of PrP was formed on refolding the disulfide-reduced protein at acidic pH, but reported that, under their conditions, the circular dichroism response interpreted as β-sheet structure was associated with protein oligomerization (45). Indeed, atomic force microscopy on oligomeric full-length β-PrP (residues 23–231) shows small, round particles, showing that it is capable of formation of oligomers without forming fibrils (35). Notably, however, salt-induced oligomeric β-PrP has been shown to be a potent inhibitor of the 26 S proteasome, in a similar manner to PrPSc (46). Impairment of the ubiquitin-proteasome system in vivo has been linked to prion neuropathology in prion-infected mice (46).Although the global properties of several PrP intermediate states have been determined (3032, 35), no information on their conformational properties on a sequence-specific basis has been obtained. Their conformational properties are considered important, as the elucidation of the chain conformation may provide information on the way in which these chains pack in the assembly process, and also potentially provide clues on the mechanism of amyloid assembly and the phenomenon of prion strains. As the conformational fluctuations and heterogeneity of molten globule states give rise to broad NMR spectra that preclude direct observation of their conformational properties by NMR (4750), here we use denaturant titration experiments to determine the conformational properties of β-PrP, through the population of the unfolded state that is visible by NMR. In addition, we use circular dichroism and analytical ultracentrifugation to examine the global structural properties, and the distribution of multimeric species that are formed from β-PrP.  相似文献   

6.
Anandamide (AEA) is an endogenous agonist of type 1 cannabinoid receptors (CB1R) that, along with metabolic enzymes of AEA and congeners, compose the “endocannabinoid system.” Here we report the biochemical, morphological, and functional characterization of the endocannabinoid system in human neuroblastoma SH-SY5Y cells that are an experimental model for neuronal cell damage and death, as well as for major human neurodegenerative disorders. We also show that AEA dose-dependently induced apoptosis of SH-SY5Y cells. Through proteomic analysis, we further demonstrate that AEA-induced apoptosis was paralleled by an ∼3 to ∼5-fold up-regulation or down-regulation of five genes; IgG heavy chain-binding protein, stress-induced phosphoprotein-1, and triose-phosphate isomerase-1, which were up-regulated, are known to act as anti-apoptotic agents; actin-related protein 2/3 complex subunit 5 and peptidylprolyl isomerase-like protein 3 isoform PPIL3b were down-regulated, and the first is required for actin network formation whereas the second is still function-orphan. Interestingly, only the effect of AEA on BiP was reversed by the CB1R antagonist SR141716, in SH-SY5Y cells as well as in human neuroblastoma LAN-5 cells (that express a functional CB1R) but not in SK-NBE cells (which do not express CB1R). Silencing or overexpression of BiP increased or reduced, respectively, AEA-induced apoptosis of SH-SY5Y cells. In addition, the expression of BiP and of the BiP-related apoptotic markers p53 and PUMA was increased by AEA through a CB1R-dependent pathway that engages p38 and p42/44 mitogen-activated protein kinases. Consistently, this effect of AEA was minimized by SR141716. In conclusion, we identified BiP as a key protein in neuronal apoptosis induced by AEA.Endocannabinoids bind to and activate both type 1 (CB1R)4 and type 2 (CB2R) cannabinoid receptors and are widely recognized as important regulators of central and peripheral functions (13). The most studied endocannabinoids are anandamide (N-arachidonoylethanolamine, AEA) and 2-arachidonoylglycerol (2-AG) (1, 3). AEA, unlike 2-AG, binds to and activates also transient receptor potential vanilloid 1 (TRPV1), thus being considered a true “endovanilloid” (4).The “endocannabinoid system (ECS)” includes the receptors, their ligands AEA and 2-AG, and the enzymes that synthesize (N-acyl-phosphatidylethanolamine-hydrolyzing phospholipase D (NAPE-PLD) and diacylglycerol lipase (DAGL)) or degrade (fatty acid amide hydrolase (FAAH) and monoacylglycerol lipase (MAGL)) AEA and 2-AG, respectively (13). On the other hand, there is still controversy about the identity and even the existence of purported “endocannabinoid transporters” that have not yet been cloned, isolated, or characterized (5). A growing interest is concerned with the ability of AEA to induce apoptosis in neuronal and non-neuronal cells (for comprehensive reviews see Refs. 69). Such a pro-apoptotic activity of AEA appears to be mediated by divergent pathways, each potentiated, reduced, or unaffected by cannabinoid or TRPV1 receptors (69). They can also be dependent on membrane cholesterol content (10, 11), generation of reactive oxygen species (12, 13), or even release of ethanolamine from AEA catalyzed by FAAH (14). In this context, it should be stressed that data on ECS and AEA-induced apoptosis in human neuronal cells are still very scant. Furthermore, a proteomic analysis of AEA-induced apoptosis has never been performed, leaving open the question of which proteins (if any) might be modulated at the level of expression upon induction of programmed cell death by this endocannabinoid. On this basis, we sought to characterize through functional and immunochemical assays and confocal microscopy the ECS components in human neuroblastoma SH-SY5Y cells. In addition, we performed a proteomic analysis of AEA-induced apoptosis in these cells that identified five proteins whose expression was modified. Among these proteins, one (BiP/GRP78) was shown to be modulated by AEA in a CB1R-dependent manner, in parallel with apoptosis. Through silencing and overexpression experiments, it was found to be a key player in the death program.  相似文献   

7.
8.
9.
10.
Fibroblasts degrade type I collagen, the major extracellular protein found in mammals, during events ranging from bulk tissue resorption to invasion through the three-dimensional extracellular matrix. Current evidence suggests that type I collagenolysis is mediated by secreted as well as membrane-anchored members of the matrix metalloproteinase (MMP) gene family. However, the roles played by these multiple and possibly redundant, degradative systems during fibroblast-mediated matrix remodeling is undefined. Herein, we use fibroblasts isolated from Mmp13−/−, Mmp8−/−, Mmp2−/−, Mmp9−/−, Mmp14−/− and Mmp16−/− mice to define the functional roles for secreted and membrane-anchored collagenases during collagen-resorptive versus collagen-invasive events. In the presence of a functional plasminogen activator-plasminogen axis, secreted collagenases arm cells with a redundant collagenolytic potential that allows fibroblasts harboring single deficiencies for either MMP-13, MMP-8, MMP-2, or MMP-9 to continue to degrade collagen comparably to wild-type fibroblasts. Likewise, Mmp14−/− or Mmp16−/− fibroblasts retain near-normal collagenolytic activity in the presence of plasminogen via the mobilization of secreted collagenases, but only Mmp14 (MT1-MMP) plays a required role in the collagenolytic processes that support fibroblast invasive activity. Furthermore, by artificially tethering a secreted collagenase to the surface of Mmp14−/− fibroblasts, we demonstrate that localized pericellular collagenolytic activity differentiates the collagen-invasive phenotype from bulk collagen degradation. Hence, whereas secreted collagenases arm fibroblasts with potent matrix-resorptive activity, only MT1-MMP confers the focal collagenolytic activity necessary for supporting the tissue-invasive phenotype.In the postnatal state, fibroblasts are normally embedded in a self-generated three-dimensional connective tissue matrix composed largely of type I collagen, the major extracellular protein found in mammals (13). Type I collagen not only acts as a structural scaffolding for the associated mesenchymal cell populations but also regulates gene expression and cell function through its interactions with collagen binding integrins and discoidin receptors (2, 4). Consistent with the central role that type I collagen plays in defining the structure and function of the extracellular matrix, the triple-helical molecule is resistant to almost all forms of proteolytic attack and can display a decades-long half-life in vivo (46). Nonetheless, fibroblasts actively remodel type I collagen during wound healing, inflammation, or neoplastic states (2, 713).To date type I collagenolytic activity is largely confined to a small subset of fewer than 10 proteases belonging to either the cysteine proteinase or matrix metalloproteinase (MMP)2 gene families (4, 1418). As all collagenases are synthesized as inactive zymogens, complex proteolytic cascades involving serine, cysteine, metallo, and aspartyl proteinases have also been linked to collagen turnover by virtue of their ability to mediate the processing of the pro-collagenases to their active forms (13, 15, 19). After activation, each collagenase can then cleave native collagen within its triple-helical domain, thus precipitating the unwinding or “melting” of the resulting collagen fragments at physiologic temperatures (4, 15). In turn, the denatured products (termed gelatin) are susceptible to further proteolysis by a broader class of “gelatinases” (4, 15). Collagen fragments are then either internalized after binding to specific receptors on the cell surface or degraded to smaller peptides with potent biological activity (2024).Previous studies by our group as well as others have identified MMPs as the primary effectors of fibroblast-mediated collagenolysis (20, 25, 26). Interestingly, adult mouse fibroblasts express at least six MMPs that can potentially degrade type I collagen, raising the possibility of multiple compensatory networks that are designed to preserve collagenolytic activity (25). Four of these collagenases belong to the family of secreted MMPs, i.e. MMP-13, MMP-8, MMP-2, and MMP-9, whereas the other two enzymes are members of the membrane-type MMP subgroup, i.e. MMP-14 (MT1-MMP) and MMP-16 (MT3-MMP) (13, 2729). From a functional perspective, the specific roles that can be assigned to secreted versus membrane-anchored collagenases remain undefined. As such, fibroblasts were isolated from either wild-type mice or mice harboring loss-of-function deletions in each of the major secreted and membrane-anchored collagenolytic genes, and the ability of the cells to degrade type I collagen was assessed. Herein, we demonstrate that fibroblasts mobilize either secreted or membrane-anchored MMPs to effectively degrade type I collagen in qualitatively and quantitatively distinct fashions. However, under conditions where fibroblasts use either secreted and membrane-anchored MMPs to exert quantitatively equivalent collagenolytic activity, only MT1-MMP plays a required role in supporting a collagen-invasive phenotype. These data establish a new paradigm wherein secreted collagenases are functionally limited to bulk collagenolytic processes, whereas MT1-MMP uniquely arms the fibroblast with a focalized degradative activity that mediates subjacent collagenolysis as well as invasion.  相似文献   

11.
Understanding the structural and assembly dynamics of the amyloid β-protein (Aβ) has direct relevance to the development of therapeutic agents for Alzheimer disease. To elucidate these dynamics, we combined scanning amino acid substitution with a method for quantitative determination of the Aβ oligomer frequency distribution, photo-induced cross-linking of unmodified proteins (PICUP), to perform “scanning PICUP.” Tyr, a reactive group in PICUP, was substituted at position 1, 10, 20, 30, or 40 (for Aβ40) or 42 (for Aβ42). The effects of these substitutions were probed using circular dichroism spectroscopy, thioflavin T binding, electron microscopy, PICUP, and mass spectrometry. All peptides displayed a random coil → α/β → β transition, but substitution-dependent alterations in assembly kinetics and conformer complexity were observed. Tyr1-substituted homologues of Aβ40 and Aβ42 assembled the slowest and yielded unusual patterns of oligomer bands in gel electrophoresis experiments, suggesting oligomer compaction had occurred. Consistent with this suggestion was the observation of relatively narrow [Tyr1]Aβ40 fibrils. Substitution of Aβ40 at the C terminus decreased the population conformational complexity and substantially extended the highest order of oligomers observed. This latter effect was observed in both Aβ40 and Aβ42 as the Tyr substitution position number increased. The ability of a single substitution (Tyr1) to alter Aβ assembly kinetics and the oligomer frequency distribution suggests that the N terminus is not a benign peptide segment, but rather that Aβ conformational dynamics and assembly are affected significantly by the competition between the N and C termini to form a stable complex with the central hydrophobic cluster.Alzheimer disease (AD)4 is the most common cause of late-life dementia (1) and is estimated to afflict more than 27 million people worldwide (2). An important etiologic hypothesis is that amyloid β-protein (Aβ) oligomers are the proximate neurotoxins in AD. Substantial in vivo and in vitro evidence supports this hypothesis (312). Neurotoxicity studies have shown that Aβ assemblies are potent neurotoxins (5, 1320), and the toxicity of some oligomers can be greater than that of the corresponding fibrils (21). Soluble Aβ oligomers inhibit hippocampal long term potentiation (4, 5, 13, 15, 17, 18, 22) and disrupt cognitive function (23). Compounds that bind and disrupt the formation of oligomers have been shown to block the neurotoxicity of Aβ (24, 25). Importantly, recent studies in higher vertebrates (dogs) have shown that substantial reduction in amyloid deposits in the absence of decreases in oligomer concentration has little effect on recovery of neurological function (26).Recent studies of Aβ oligomers have sought to correlate oligomer size and biological activity. Oligomers in the supernatants of fibril preparations centrifuged at 100,000 × g caused sustained calcium influx in rat hippocampal neurons, leading to calpain activation and dynamin 1 degradation (27). Aβ-derived diffusible ligand-like Aβ42 oligomers induced inflammatory responses in cultured rat astrocytes (28). A 90-kDa Aβ42 oligomer (29) has been shown to activate ERK1/2 in rat hippocampal slices (30) and bind avidly to human cortical neurons (31), in both cases causing apoptotic cell death. A comparison of the time dependence of the toxic effects of the 90-kDa assembly with that of Aβ-derived diffusible ligands revealed a 5-fold difference, Aβ-derived diffusible ligands requiring more time for equivalent effects (31). A 56-kDa oligomer, “Aβ*56,” was reported to cause memory impairment in middle-aged transgenic mice expressing human amyloid precursor protein (32). A nonamer also had adverse effects. Impaired long term potentiation in rat brain slices has been attributed to Aβ trimers identified in media from cultured cells expressing human amyloid precursor protein (33). Dimers and trimers from this medium also have been found to cause progressive loss of synapses in organotypic rat hippocampal slices (10). In mice deficient in neprilysin, an enzyme that has been shown to degrade Aβ in vivo (34), impairment in neuronal plasticity and cognitive function correlated with significant increases in Aβ dimer levels and synapse-associated Aβ oligomers (35).The potent pathologic effects of Aβ oligomers provide a compelling reason for elucidating the mechanism(s) of their formation. This has been a difficult task because of the metastability and polydispersity of Aβ assemblies (36). To obviate these problems, we introduced the use of the method of photo-induced cross-linking of unmodified proteins (PICUP) to rapidly (<1 s) and covalently stabilize oligomer mixtures (for reviews see Refs. 37, 38). Oligomers thus stabilized no longer exist in equilibrium with monomers or each other, allowing determination of oligomer frequency distributions by simple techniques such as SDS-PAGE (37). Recently, to obtain population-average information on contributions to fibril formation of amino acid residues at specific sites in Aβ, we employed a scanning intrinsic fluorescence approach (39). Tyr was used because it is a relatively small fluorophore, exists natively in Aβ, and possesses the side chain most reactive in the PICUP chemistry (40). Using this approach, we found that the central hydrophobic cluster region (Leu17–Ala21) was particularly important in controlling fibril formation of Aβ40, whereas the C terminus was the predominant structural element controlling Aβ42 assembly (39). Here we present results of studies in which key strategic features of the two methods have been combined to enable execution of “scanning PICUP” and the consequent revelation of site-specific effects on Aβ oligomerization.  相似文献   

12.
Early onset generalized dystonia (DYT1) is an autosomal dominant neurological disorder caused by deletion of a single glutamate residue (torsinA ΔE) in the C-terminal region of the AAA+ (ATPases associated with a variety of cellular activities) protein torsinA. The pathogenic mechanism by which torsinA ΔE mutation leads to dystonia remains unknown. Here we report the identification and characterization of a 628-amino acid novel protein, printor, that interacts with torsinA. Printor co-distributes with torsinA in multiple brain regions and co-localizes with torsinA in the endoplasmic reticulum. Interestingly, printor selectively binds to the ATP-free form but not to the ATP-bound form of torsinA, supporting a role for printor as a cofactor rather than a substrate of torsinA. The interaction of printor with torsinA is completely abolished by the dystonia-associated torsinA ΔE mutation. Our findings suggest that printor is a new component of the DYT1 pathogenic pathway and provide a potential molecular target for therapeutic intervention in dystonia.Early onset generalized torsion dystonia (DYT1) is the most common and severe form of hereditary dystonia, a movement disorder characterized by involuntary movements and sustained muscle spasms (1). This autosomal dominant disease has childhood onset and its dystonic symptoms are thought to result from neuronal dysfunction rather than neurodegeneration (2, 3). Most DYT1 cases are caused by deletion of a single glutamate residue at positions 302 or 303 (torsinA ΔE) of the 332-amino acid protein torsinA (4). In addition, a different torsinA mutation that deletes amino acids Phe323–Tyr328 (torsinA Δ323–328) was identified in a single family with dystonia (5), although the pathogenic significance of this torsinA mutation is unclear because these patients contain a concomitant mutation in another dystonia-related protein, ϵ-sarcoglycan (6). Recently, genetic association studies have implicated polymorphisms in the torsinA gene as a genetic risk factor in the development of adult-onset idiopathic dystonia (7, 8).TorsinA contains an N-terminal endoplasmic reticulum (ER)3 signal sequence and a 20-amino acid hydrophobic region followed by a conserved AAA+ (ATPases associated with a variety of cellular activities) domain (9, 10). Because members of the AAA+ family are known to facilitate conformational changes in target proteins (11, 12), it has been proposed that torsinA may function as a molecular chaperone (13, 14). TorsinA is widely expressed in brain and multiple other tissues (15) and is primarily associated with the ER and nuclear envelope (NE) compartments in cells (1620). TorsinA is believed to mainly reside in the lumen of the ER and NE (1719) and has been shown to bind lamina-associated polypeptide 1 (LAP1) (21), lumenal domain-like LAP1 (LULL1) (21), and nesprins (22). In addition, recent evidence indicates that a significant pool of torsinA exhibits a topology in which the AAA+ domain faces the cytoplasm (20). In support of this topology, torsinA is found in the cytoplasm, neuronal processes, and synaptic terminals (2, 3, 15, 2326) and has been shown to bind cytosolic proteins snapin (27) and kinesin light chain 1 (20). TorsinA has been proposed to play a role in several cellular processes, including dopaminergic neurotransmission (2831), NE organization and dynamics (17, 22, 32), and protein trafficking (27, 33). However, the precise biological function of torsinA and its regulation remain unknown.To gain insights into torsinA function, we performed yeast two-hybrid screens to search for torsinA-interacting proteins in the brain. We report here the isolation and characterization of a novel protein named printor (protein interactor of torsinA) that interacts selectively with wild-type (WT) torsinA but not the dystonia-associated torsinA ΔE mutant. Our data suggest that printor may serve as a cofactor of torsinA and provide a new molecular target for understanding and treating dystonia.  相似文献   

13.
Myofilament proteins are responsible for cardiac contraction. The myofilament subproteome, however, has not been comprehensively analyzed thus far. In the present study, cardiomyocytes were isolated from rodent hearts and stimulated with endothelin-1 and isoproterenol, potent inducers of myofilament protein phosphorylation. Subsequently, cardiomyocytes were “skinned,” and the myofilament subproteome was analyzed using a high mass accuracy ion trap tandem mass spectrometer (LTQ Orbitrap XL) equipped with electron transfer dissociation. As expected, a small number of myofilament proteins constituted the majority of the total protein mass with several known phosphorylation sites confirmed by electron transfer dissociation. More than 600 additional proteins were identified in the cardiac myofilament subproteome, including kinases and phosphatase subunits. The proteomic comparison of myofilaments from control and treated cardiomyocytes suggested that isoproterenol treatment altered the subcellular localization of protein phosphatase 2A regulatory subunit B56α. Immunoblot analysis of myocyte fractions confirmed that β-adrenergic stimulation by isoproterenol decreased the B56α content of the myofilament fraction in the absence of significant changes for the myosin phosphatase target subunit isoforms 1 and 2 (MYPT1 and MYPT2). Furthermore, immunolabeling and confocal microscopy revealed the spatial redistribution of these proteins with a loss of B56α from Z-disc and M-band regions but increased association of MYPT1/2 with A-band regions of the sarcomere following β-adrenergic stimulation. In summary, we present the first comprehensive proteomics data set of skinned cardiomyocytes and demonstrate the potential of proteomics to unravel dynamic changes in protein composition that may contribute to the neurohormonal regulation of myofilament contraction.Myofilament proteins comprise the fundamental contractile apparatus of the heart, the cardiac sarcomere. They are subdivided into thin filament proteins, including actin, tropomyosin, the troponin complex (troponin C, troponin I, and troponin T), and thick filament proteins, including myosin heavy chains, myosin light chains, and myosin-binding protein C. Although calcium is the principal regulator of cardiac contraction through the excitation-contraction coupling process that culminates in calcium binding to troponin C, myofilament function is also significantly modulated by phosphorylation of constituent proteins, such as cardiac troponin I (cTnI),1 cardiac myosin-binding protein C (cMyBP-C), and myosin regulatory light chain (MLC-2). “Skinned” myocyte preparations from rodent hearts, in which the sarcolemmal envelope is disrupted through the use of detergents, have been invaluable in providing mechanistic information on the functional consequences of myofilament protein phosphorylation following exposure to neurohormonal stimuli that activate pertinent kinases prior to skinning or direct exposure to such kinases in active form after skinning (for recent examples, see studies on the phosphorylation of cTnI (13), cMyBP-C (46), and MLC-2 (79)). Nevertheless, to date, only a few myofilament proteins have been studied using proteomics (1019), and a detailed proteomic characterization of the myofilament subproteome and its associated proteins from skinned myocytes has not been performed. In the present analysis, we used an LTQ Orbitrap XL equipped with ETD (20) to analyze the subproteome of skinned cardiomyocytes with or without prior stimulation. Endothelin-1 and isoproterenol were used to activate the endothelin receptor/protein kinase C and β-adrenoreceptor/protein kinase A pathway, respectively (21, 22). Importantly, the mass accuracy of the Orbitrap mass analyzer helped to distinguish true phosphorylation sites from false assignments, and the sensitivity of the ion trap provided novel insights into the translocation of phosphatase regulatory and targeting subunits following β-adrenergic stimulation.  相似文献   

14.
15.
16.
17.
Clinically, amniotic membrane (AM) suppresses inflammation, scarring, and angiogenesis. AM contains abundant hyaluronan (HA) but its function in exerting these therapeutic actions remains unclear. Herein, AM was extracted sequentially with buffers A, B, and C, or separately by phosphate-buffered saline (PBS) alone. Agarose gel electrophoresis showed that high molecular weight (HMW) HA (an average of ∼3000 kDa) was predominantly extracted in isotonic Extract A (70.1 ± 6.0%) and PBS (37.7 ± 3.2%). Western blot analysis of these extracts with hyaluronidase digestion or NaOH treatment revealed that HMW HA was covalently linked with the heavy chains (HCs) of inter-α-inhibitor (IαI) via a NaOH-sensitive bond, likely transferred by the tumor necrosis factor-α stimulated gene-6 protein (TSG-6). This HC·HA complex (nHC·HA) could be purified from Extract PBS by two rounds of CsCl/guanidine HCl ultracentrifugation as well as in vitro reconstituted (rcHC·HA) by mixing HMW HA, serum IαI, and recombinant TSG-6. Consistent with previous reports, Extract PBS suppressed transforming growth factor-β1 promoter activation in corneal fibroblasts and induced mac ro phage apo pto sis. However, these effects were abolished by hyaluronidase digestion or heat treatment. More importantly, the effects were retained in the nHC·HA or rcHC·HA. These data collectively suggest that the HC·HA complex is the active component in AM responsible in part for clinically observed anti-inflammatory and anti-scarring actions.Hyaluronan (HA)4 is widely distributed in extracellular matrices, tissues, body fluids, and even in intracellular compartments (reviewed in Refs. 1 and 2). The molecular weight of HA ranges from 200 to 10,000 kDa depending on the source (3), but can also exist as smaller fragments and oligosaccharides under certain physiological or pathological conditions (1). Investigations over the last 15 years have suggested that low Mr HA can induce the gene expression of proinflammatory mediators and proangiogenesis, whereas high molecular weight (HMW) HA inhibits these processes (47).Several proteins have been shown to bind to HA (8) such as aggrecan (9), cartilage link protein (10), versican (11), CD44 (12, 13), inter-α-inhibitor (IαI) (14, 15), and tumor necrosis factor-α stimulated gene-6 protein (TSG-6) (16, 17). IαI consists of two heavy chains (HCs) (HC1 and HC2), both of which are linked through ester bonds to a chondroitin sulfate chain that is attached to the light chain, i.e. bikunin. Among all HA-binding proteins, only the HCs of IαI have been clearly demonstrated to be covalently coupled to HA (14, 18). However, TSG-6 has also been reported to form stable, possibly covalent, complexes with HA, either alone (19, 20) or when associated with HC (21).The formation of covalent bonds between HCs and HA is mediated by TSG-6 (2224) where its expression is often induced by inflammatory mediators such as tumor necrosis factor-α and interleukin-1 (25, 26). TSG-6 is also expressed in inflammatory-like processes, such as ovulation (21, 27, 28) and cervical ripening (29). TSG-6 interacts with both HA (17) and IαI (21, 24, 3033), and is essential for covalently transferring HCs on to HA (2224). The TSG-6-mediated formation of the HC·HA complex has been demonstrated to play a crucial role in female fertility in mice. The HC·HA complex is an integral part of an expanded extracellular “cumulus” matrix around the oocyte, which plays a critical role in successful ovulation and fertilization in vivo (22, 34). HC·HA complexes have also been found at sites of inflammation (3538) where its pro- or anti-inflammatory role remain arguable (39, 40).Immunostaining reveals abundant HA in the avascular stromal matrix of the AM (41, 42).5 In ophthalmology, cryopreserved AM has been widely used as a surgical graft for ocular surface reconstruction and exerts clinically observable actions to promote epithelial wound healing and to suppress inflammation, scarring, and angiogenesis (for reviews see Refs. 4345). However, it is not clear whether HA in AM forms HC·HA complex, and if so whether such an HC·HA complex exerts any of the above therapeutic actions. To address these questions, we extracted AM with buffers of increasing salt concentration. Because HMW HA was found to form the HC·HA complex and was mainly extractable by isotonic solutions, we further purified it from the isotonic AM extract and reconstituted it in vitro from three defined components, i.e. HMW HA, serum IαI, and recombinant TSG-6. Our results showed that the HC·HA complex is an active component in AM responsible for the suppression of TGF-β1 promoter activity, linkable to the scarring process noted before by AM (4648) and by the AM soluble extract (49), as well as for the promotion of macrophage death, linkable to the inflammatory process noted by AM (50) and the AM soluble extract (51).  相似文献   

18.
19.
20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

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