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.
STIM1 and ORAI1, the two limiting components in the Ca2+ release-activated Ca2+ (CRAC) signaling cascade, have been reported to interact upon store depletion, culminating in CRAC current activation. We have recently identified a modulatory domain between amino acids 474 and 485 in the cytosolic part of STIM1 that comprises 7 negatively charged residues. A STIM1 C-terminal fragment lacking this domain exhibits enhanced interaction with ORAI1 and 2–3-fold higher ORAI1/CRAC current densities. Here we focused on the role of this CRAC modulatory domain (CMD) in the fast inactivation of ORAI1/CRAC channels, utilizing the whole-cell patch clamp technique. STIM1 mutants either with C-terminal deletions including CMD or with 7 alanines replacing the negative amino acids within CMD gave rise to ORAI1 currents that displayed significantly reduced or even abolished inactivation when compared with STIM1 mutants with preserved CMD. Consistent results were obtained with cytosolic C-terminal fragments of STIM1, both in ORAI1-expressing HEK 293 cells and in RBL-2H3 mast cells containing endogenous CRAC channels. Inactivation of the latter, however, was much more pronounced than that of ORAI1. The extent of inactivation of ORAI3 channels, which is also considerably more prominent than that of ORAI1, was also substantially reduced by co-expression of STIM1 constructs missing CMD. Regarding the dependence of inactivation on Ca2+, a decrease in intracellular Ca2+ chelator concentrations promoted ORAI1 current fast inactivation, whereas Ba2+ substitution for extracellular Ca2+ completely abrogated it. In summary, CMD within the STIM1 cytosolic part provides a negative feedback signal to Ca2+ entry by triggering fast Ca2+-dependent inactivation of ORAI/CRAC channels.The Ca2+ release-activated Ca2+ (CRAC)5 channel is one of the best characterized store-operated entry pathways (17). Substantial efforts have led to identification of two key components of the CRAC channel machinery: the stromal interaction molecule 1 (STIM1), which is located in the endoplasmic reticulum and acts as a Ca2+ sensor (810), and ORAI1/CRACM1, the pore-forming subunit of the CRAC channel (1113). Besides ORAI1, two further homologues named ORAI2 and ORAI3 belong to the ORAI channel family (12, 14).STIM1 senses endoplasmic reticulum store depletion primarily by its luminal EF-hand in its N terminus (8, 15), redistributes close to the plasma membrane, where it forms puncta-like structures, and co-clusters with ORAI1, leading to inward Ca2+ currents (12, 1619). The STIM1 C terminus, located in the cytosol, contains two coiled-coil regions overlapping with an ezrin-radixin-moesin (ERM)-like domain followed by a serine/proline- and a lysine-rich region (2, 8, 2022). Three recent studies have described the essential ORAI-activating region within the ERM domain, termed SOAR (Stim ORAI-activating region) (23), OASF (ORAI-activating small fragment) (24), and CAD (CRAC-activating domain) (25), including the second coiled coil domain and the following ∼55 amino acids. We and others have provided evidence that store depletion leads to a dynamic coupling of STIM1 to ORAI1 (2628) that is mediated by a direct interaction of the STIM1 C terminus with ORAI1 C terminus probably involving the putative coiled-coil domain in the latter (27).Furthermore, different groups have proven that the C terminus of STIM1 is sufficient to activate CRAC as well as ORAI1 channels independent of store depletion (2225, 27, 29). We have identified that OASF-(233–474) or shorter fragments exhibit further enhanced coupling to ORAI1 resulting in 3-fold increased constitutive Ca2+ currents. A STIM1 fragment containing an additional cluster of anionic amino acids C-terminal to position 474 displays weaker interaction with ORAI1 as well as reduced Ca2+ current comparable with that mediated by wild-type STIM1 C terminus. Hence, we have suggested that these 11 amino acids (474–485) act in a modulatory manner onto ORAI1; however, their detailed mechanistic impact within the STIM1/ORAI1 signaling machinery has remained so far unclear.In this study, we focused on the impact of this negative cluster on fast inactivation of STIM1-mediated ORAI Ca2+ currents. Lis et al. (30) have shown that all three ORAI homologues display distinct inactivation profiles, where ORAI2 and ORAI3 show a much more pronounced fast inactivation than ORAI1. Moreover, it has been reported (31) that different expression levels of STIM1 to ORAI1 affect the properties of CRAC current inactivation. Yamashita et al. (32) have demonstrated a linkage between the selectivity filter of ORAI1 and its Ca2+-dependent fast inactivation. Here we provide evidence that a cluster of acidic residues within the C terminus of STIM1 is involved in the fast inactivation of ORAI1 and further promotes that of ORAI3 and native CRAC currents.  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.
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   

4.
SLC26A7 (human)/Slc26a7 (mouse) is a recently identified chloride-base exchanger and/or chloride transporter that is expressed on the basolateral membrane of acid-secreting cells in the renal outer medullary collecting duct (OMCD) and in gastric parietal cells. Here, we show that mice with genetic deletion of Slc26a7 expression develop distal renal tubular acidosis, as manifested by metabolic acidosis and alkaline urine pH. In the kidney, basolateral Cl/HCO3 exchange activity in acid-secreting intercalated cells in the OMCD was significantly decreased in hypertonic medium (a normal milieu for the medulla) but was reduced only mildly in isotonic medium. Changing from a hypertonic to isotonic medium (relative hypotonicity) decreased the membrane abundance of Slc26a7 in kidney cells in vivo and in vitro. In the stomach, stimulated acid secretion was significantly impaired in isolated gastric mucosa and in the intact organ. We propose that SLC26A7 dysfunction should be investigated as a potential cause of unexplained distal renal tubular acidosis or decreased gastric acid secretion in humans.The collecting duct segment of the distal kidney nephron plays a major role in systemic acid base homeostasis by acid secretion and bicarbonate absorption. The acid secretion occurs via H+-ATPase and H-K-ATPase into the lumen and bicarbonate is absorbed via basolateral Cl/HCO3 exchangers (14). The tubules, which are located within the outer medullary region of the kidney collecting duct (OMCD),2 have the highest rate of acid secretion among the distal tubule segments and are therefore essential to the maintenance of acid base balance (2).The gastric parietal cell is the site of generation of acid and bicarbonate through the action of cytosolic carbonic anhydrase II (5, 6). The intracellular acid is secreted into the lumen via gastric H-K-ATPase, which works in conjunction with a chloride channel and a K+ recycling pathway (710). The intracellular bicarbonate is transported to the blood via basolateral Cl/HCO3 exchangers (1114).SLC26 (human)/Slc26 (mouse) isoforms are members of a conserved family of anion transporters that display tissue-specific patterns of expression in epithelial cells (1524). Several SLC26 members can function as chloride/bicarbonate exchangers. These include SLC26A3 (DRA), SLC26A4 (pendrin), SLC26A6 (PAT1 or CFEX), SLC26A7, and SLC26A9 (2531). SLC26A7 and SLC26A9 can also function as chloride channels (3234).SLC26A7/Slc26a7 is predominantly expressed in the kidney and stomach (28, 29). In the kidney, Slc26a7 co-localizes with AE1, a well-known Cl/HCO3 exchanger, on the basolateral membrane of (acid-secreting) A-intercalated cells in OMCD cells (29, 35, 36) (supplemental Fig. 1). In the stomach, Slc26a7 co-localizes with AE2, a major Cl/HCO3 exchanger, on the basolateral membrane of acid secreting parietal cells (28). To address the physiological function of Slc26a7 in the intact mouse, we have generated Slc26a7 ko mice. We report here that Slc26a7 ko mice exhibit distal renal tubular acidosis and impaired gastric acidification in the absence of morphological abnormalities in kidney or stomach.  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.
7.
8.
9.
10.
Molecular identification of the Ca2+-dependent chloride channel TMEM16A (ANO1) provided a fundamental step in understanding Ca2+-dependent Cl secretion in epithelia. TMEM16A is an intrinsic constituent of Ca2+-dependent Cl channels in cultured epithelia and may control salivary output, but its physiological role in native epithelial tissues remains largely obscure. Here, we demonstrate that Cl secretion in native epithelia activated by Ca2+-dependent agonists is missing in mice lacking expression of TMEM16A. Ca2+-dependent Cl transport was missing or largely reduced in isolated tracheal and colonic epithelia, as well as hepatocytes and acinar cells from pancreatic and submandibular glands of TMEM16A−/− animals. Measurement of particle transport on the surface of tracheas ex vivo indicated largely reduced mucociliary clearance in TMEM16A−/− mice. These results clearly demonstrate the broad physiological role of TMEM16A−/− for Ca2+-dependent Cl secretion and provide the basis for novel treatments in cystic fibrosis, infectious diarrhea, and Sjöegren syndrome.Electrolyte secretion in epithelial tissues is based on the major second messenger pathways cAMP and Ca2+, which activate the cystic fibrosis transmembrane conductance regulator (CFTR)2 Cl channels and Ca2+-dependent Cl channels, respectively (13). CFTR conducts Cl in epithelial cells of airways, intestine, and the ducts of pancreas and sweat gland, while Ca2+-dependent Cl channels secrete Cl in pancreatic acini and salivary and sweat glands (46). Controversy exists as to the contribution of these channels to Cl secretion in submucosal glands of airways and the relevance for cystic fibrosis (79). While cAMP-dependent Cl secretion by CFTR is well examined, detailed analysis of epithelial Ca2+-dependent Cl secretion is hampered by the lack of a molecular counterpart. Although bestrophins may form Ca2+-dependent Cl channels and facilitate Ca2+-dependent Cl secretion in epithelial tissues (10, 11), they are unlikely to form secretory Cl channels in the apical cell membrane, because Ca2+-dependent Cl secretion is still present in epithelia of mice lacking expression of bestrophin (12). Bestrophins may rather have an intracellular function by facilitating receptor mediated Ca2+ signaling and activation of membrane localized channels (13). With the discovery that TMEM16A produces Ca2+-activated Cl currents with biophysical and pharmacological properties close to those in native epithelial tissues, these proteins are now very likely candidates for endogenous Ca2+-dependent Cl channels (1417). In cultured airway epithelial cells, small interfering RNA knockdown of endogenous TMEM16A largely reduced calcium-dependent chloride secretion (16). However, apart from preliminary studies of airways and salivary glands, the physiological significance of TMEM16A in native epithelia, particularly in glands, is unclear (14, 17).  相似文献   

11.
12.
Collapsin response mediator protein 2 (CRMP2) is an intracellular protein that mediates signaling of Semaphorin3A (Sema3A), a repulsive axon guidance molecule. Fyn, a Src-type tyrosine kinase, is involved in the Sema3A signaling. However, the relationship between CRMP2 and Fyn in this signaling pathway is still unknown. In our research, we demonstrated that Fyn phosphorylated CRMP2 at Tyr32 residues in HEK293T cells. Immunohistochemical analysis using a phospho-specific antibody at Tyr32 of CRMP showed that Tyr32-phosphorylated CRMP was abundant in the nervous system, including dorsal root ganglion neurons, the molecular and Purkinje cell layer of adult cerebellum, and hippocampal fimbria. Overexpression of a nonphosphorylated mutant (Tyr32 to Phe32) of CRMP2 in dorsal root ganglion neurons interfered with Sema3A-induced growth cone collapse response. These results suggest that Fyn-dependent phosphorylation of CRMP2 at Tyr32 is involved in Sema3A signaling.Collapsin response mediator proteins (CRMPs)4 have been identified as intracellular proteins that mediate Semaphorin3A (Sema3A) signaling in the nervous system (1). CRMP2 is one of the five members of the CRMP family. CRMPs also mediate signal transduction of NT3, Ephrin, and Reelin (24). CRMPs interact with several intracellular molecules, including tubulin, Numb, kinesin1, and Sra1 (58). CRMPs are involved in axon guidance, axonal elongation, cell migration, synapse maturation, and the generation of neuronal polarity (1, 2, 4, 5).CRMP family proteins are known to be the major phosphoproteins in the developing brain (1, 9). CRMP2 is phosphorylated by several Ser/Thr kinases, such as Rho kinase, cyclin-dependent kinase 5 (Cdk5), and glycogen synthase kinase 3β (GSK3β) (2, 1013). The phosphorylation sites of CRMP2 by these kinases are clustered in the C terminus and have already been identified. Rho kinase phosphorylates CRMP2 at Thr555 (10). Cdk5 phosphorylates CRMP2 at Ser522, and this phosphorylation is essential for sequential phosphorylations by GSK3β at Ser518, Thr514, and Thr509 (2, 1113). These phosphorylations disrupt the interaction of CRMP2 with tubulin or Numb (2, 3, 13). The sequential phosphorylation of CRMP2 by Cdk5 and GSK3β is an essential step in Sema3A signaling (11, 13). Furthermore, the neurofibrillary tangles in the brains of people with Alzheimer disease contain hyperphosphorylated CRMP2 at Thr509, Ser518, and Ser522 (14, 15).CRMPs are also substrates of several tyrosine kinases. The phosphorylation of CRMP2 by Fes/Fps and Fer has been shown to be involved in Sema3A signaling (16, 17). Phosphorylation of CRMP2 at Tyr479 by a Src family tyrosine kinase Yes regulates CXCL12-induced T lymphocyte migration (18). We reported previously that Fyn is involved in Sema3A signaling (19). Fyn associates with PlexinA2, one of the components of the Sema3A receptor complex. Fyn also activates Cdk5 through the phosphorylation at Tyr15 of Cdk5 (19). In dorsal root ganglion (DRG) neurons from fyn-deficient mice, Sema3A-induced growth cone collapse response is attenuated compared with control mice (19). Furthermore, we recently found that Fyn phosphorylates CRMP1 and that this phosphorylation is involved in Reelin signaling (4). Although it has been shown that CRMP2 is involved in Sema3A signaling (1, 11, 13), the relationship between Fyn and CRMP2 in Sema3A signaling and the tyrosine phosphorylation site(s) of CRMPs remain unknown.Here, we show that Fyn phosphorylates CRMP2 at Tyr32. Using a phospho-specific antibody against Tyr32, we determined that the residue is phosphorylated in vivo. A nonphosphorylated mutant CRMP2Y32F inhibits Sema3A-induced growth cone collapse. These results indicate that tyrosine phosphorylation by Fyn at Tyr32 is involved in Sema3A signaling.  相似文献   

13.
The reversible regulation of myosin light chain phosphatase (MLCP) in response to agonist stimulation and cAMP/cGMP signals plays an important role in the regulation of smooth muscle (SM) tone. Here, we investigated the mechanism underlying the inhibition of MLCP induced by the phosphorylation of myosin phosphatase targeting subunit (MYPT1), a regulatory subunit of MLCP, at Thr-696 and Thr-853 using glutathione S-transferase (GST)-MYPT1 fragments having the inhibitory phosphorylation sites. GST-MYPT1 fragments, including only Thr-696 and only Thr-853, inhibited purified MLCP (IC50 = 1.6 and 60 nm, respectively) when they were phosphorylated with RhoA-dependent kinase (ROCK). The activities of isolated catalytic subunits of type 1 and type 2A phosphatases (PP1 and PP2A) were insensitive to either fragment. Phospho-GST-MYPT1 fragments docked directly at the active site of MLCP, and this was blocked by a PP1/PP2A inhibitor microcystin (MC)-LR or by mutation of the active sites in PP1. GST-MYPT1 fragments induced a contraction of β-escin-permeabilized ileum SM at constant pCa 6.3 (EC50 = 2 μm), which was eliminated by Ala substitution of the fragment at Thr-696 or by ROCK inhibitors or 8Br-cGMP. GST-MYPT1-(697–880) was 5-times less potent than fragments including Thr-696. Relaxation induced by 8Br-cGMP was not affected by Ala substitution at Ser-695, a known phosphorylation site for protein kinase A/G. Thus, GST-MYPT1 fragments are phosphorylated by ROCK in permeabilized SM and mimic agonist-induced inhibition and cGMP-induced activation of MLCP. We propose a model in which MYPT1 phosphorylation at Thr-696 and Thr-853 causes an autoinhibition of MLCP that accounts for Ca2+ sensitization of smooth muscle force.The contractile state of smooth muscle (SM)3 is driven by phosphorylation of the regulatory myosin light chain and reflects the balance of the Ca2+-calmodulin-dependent myosin light chain kinase and myosin light chain phosphatase (MLCP) activities (1). The stoichiometry between force and [Ca2+] varies with different agonists (2), reflecting other signaling pathways that modulate the MLCP or myosin light chain kinase activities (35). Agonist activation of G-protein-coupled receptors triggers Ca2+ release from the sarcoplasmic reticulum. Simultaneously, G-protein-coupled receptor signals are mediated by Ca2+-independent phospholipase A2 (6) and initiate kinase signals, such as PKC, phosphoinositide 3-kinase (7), and ROCK. These lead to inhibition of MLCP activity resulting in an increase in regulatory myosin light chain phosphorylation independent of a change in Ca2+ (Ca2+ sensitization) (for review, see Ref. 1). K+ depolarization can also activate RhoA in a Ca2+-dependent manner (8). Conversely, Ca2+ desensitization occurs when nitric oxide production and the activation of Gas elevate cGMP and cAMP levels in SM, leading to dis-inhibition and restoration of MLCP activity (915). Thus, MLCP plays a pivotal role in controlling phosphorylation of myosin, in response to physiological stimulation.MLCP is a trimeric holoenzyme consisting of a catalytic subunit of protein phosphatase 1 (PP1) δ isoform and a regulatory complex of MYPT1 and an accessory M21 subunit (16). A PP1 binding site, KVKF38, is located at the N terminus of MYPT1 followed by an ankyrin-repeat domain. This N-terminal domain forms a part of the active site together with the catalytic subunit and controls the substrate specificity via allosteric interaction and targeting to loci (17). The C-terminal region of MYPT1 directly binds to substrates such as myosin and ezrin/radixin/moecin proteins as well as, under some conditions, the plasma membrane, tethering the catalytic subunit to multiple targets (18, 19). Furthermore, MYPT1 is involved in the regulation of MLCP activity. Alternative splicing of MYPT1 occurs in SM depending on the tissue and the developmental stage (20). An exon 13 splicing of MYPT1 is involved in Ca2+ sensitization that occurs in response to GTP (21), whereas a splice variant of MYPT1, containing the C-terminal Leu-zipper sequence, correlates with cGMP-dependent relaxation of smooth muscle (22). Direct binding of PKG to MYPT1 at the Leu-zipper domain and/or Arg/Lys-rich domain is involved in the activation of MLCP (2325). In addition, a myosin phosphatase-Rho interacting protein (M-RIP) is directly associated with the MYPT1 C-terminal domain, proposed to recruit RhoA to the MLCP complex (26). The C-terminal region also binds to ZIP kinase, which phosphorylates MYPT1 at Thr-6964 (27). Thus, the C-terminal domain of MYPT1 functions as a scaffold for multiple phosphatase regulatory proteins.Phosphorylation of MYPT1 at Thr-696 and Thr-853 and the phosphatase inhibitory protein CPI-17 at Thr-38 play dominant roles in the agonist-induced inhibition of MLCP (18, 2834), yet the molecular mechanism(s) of MYPT1 inhibitory phosphorylation is poorly understood. Receptor activation induces biphasic contraction of SM, reflecting a sequential activation of PKC and ROCK. Phosphorylation of CPI-17 occurs first in parallel with Ca2+ release and the activation of a conventional PKC that causes Ca2+-dependent Ca2+ sensitization (35). A delayed activation of ROCK increases the phosphorylation of MYPT1 at Thr-853. These phosphorylation events maintain the sustained phase of contraction after the fall in [Ca2+]i (35). Phosphorylation of MYPT1 at Thr-853 is elevated in response to various agonists (35, 36). Unlike the Thr-853 site, phosphorylation of MYPT1 at Thr-696 is often spontaneously phosphorylated under resting conditions and insensitive to stimuli with most agonists (36). Nonetheless, up-regulation of MYPT1 phosphorylation at Thr-696 is reported in some types of hypertensive animals and patients, suggesting an importance of the site under pathological conditions (3739). Phosphorylation of CPI-17 and MYPT1 at Thr-696 is reversed in response to nitric oxide production and cGMP elevation, which parallels relaxation (14, 15). Upon cGMP elevation, MYPT1 at Ser-695 is phosphorylated, and the Ser phosphorylation blocks the adjacent phosphorylation at Thr-696, causing dis-inhibition of MLCP (27, 40). However, Ser-695 phosphorylation does not cause the dephosphorylation at Thr-696 in intact cerebral artery (41). Thus, phosphorylation of MYPT1 governs Ca2+ sensitization and desensitization of SM, although the underlying mechanisms are still controversial. In addition, telokin, a dominant protein in visceral and phasic vascular SM tissues, is phosphorylated by PKG and PKA, activating MLCP by an unknown mechanism and inducing SM relaxation (42).Multiple mechanisms have been suggested for the phosphorylation-dependent inhibition of MLCP. Thiophosphorylation of MYPT1 results in lower Vm and higher Km values of MLCP activity, suggesting that allosteric modulation of the active site is necessary for the thiophosphorylation-dependent inhibition of MLCP (43). On the other hand, translocation of MYPT1 to the plasma membrane region occurs in parallel with the phosphorylation of MYPT1 at Thr-696 (44, 45), but the amount translocated and the functional meaning remain controversial (41). Phosphorylation of MYPT1 at Thr-853 in vitro reduces its affinity for phospho-myosin, thus suppressing the phosphatase activity (18). It has also been demonstrated that reconstitution of thiophosphorylated MYPT1 at Thr-696 or Thr-853 with isolated PP1δ produces a less-active form of MLCP complex (46). This supports the kinetic analysis (43) that suggests an allosteric effect of MYPT1 phosphorylation on the phosphatase activity. In contrast, a thiophosphopeptide mimicking the phosphorylation site of MBS85, a homolog of MYPT1 and not present in SM, inhibits the activity of MBS85·PP1 complex, suggesting the direct interaction between the MBS85 site and PP1 (47). In the crystal structure model of MYPT1-(1–229). PP1δ complex, the electrostatic potential map at the MLCP active site complements amino acid profiles around the phosphorylation sites (17). Therefore, it is possible that the inhibitory phosphorylation sites directly dock at the active site of MLCP and inhibit the activity. Here, we examine mechanisms underlying the inhibition of MLCP through the phosphorylation of MYPT1 at Thr-696 and Thr-853 using GST fusion versions of various MYPT1 fragments including or excluding either or both of these phosphorylation sites. Phosphorylated MYPT1 fragments including either Thr-696 or Thr-853 potently and specifically inhibit MLCP purified from pig aorta and the enzyme associated with myofilaments in permeabilized ileum SM tissues. We further show that inhibition of MLCP in SM tissues is eliminated by activation of PKA/PKG, suggesting that the GST-MYPT1 fragments mimic agonist-induced autoinhibition and cAMP/cGMP-dependent dis-autoinhibition of MLCP in SM.  相似文献   

14.
We have investigated in detail the role of intra-organelle Ca2+ content during induction of apoptosis by the oxidant menadione while changing and monitoring the Ca2+ load of endoplasmic reticulum (ER), mitochondria, and acidic organelles. Menadione causes production of reactive oxygen species, induction of oxidative stress, and subsequently apoptosis. In both pancreatic acinar and pancreatic tumor AR42J cells, menadione was found to induce repetitive cytosolic Ca2+ responses because of the release of Ca2+ from both ER and acidic stores. Ca2+ responses to menadione were accompanied by elevation of Ca2+ in mitochondria, mitochondrial depolarization, and mitochondrial permeability transition pore (mPTP) opening. Emptying of both the ER and acidic Ca2+ stores did not necessarily prevent menadione-induced apoptosis. High mitochondrial Ca2+ at the time of menadione application was the major factor determining cell fate. However, if mitochondria were prevented from loading with Ca2+ with 10 μm RU360, then caspase-9 activation did not occur irrespective of the content of other Ca2+ stores. These results were confirmed by ratiometric measurements of intramitochondrial Ca2+ with pericam. We conclude that elevated Ca2+ in mitochondria is the crucial factor in determining whether cells undergo oxidative stress-induced apoptosis.Apoptosis, a mechanism of programmed cell death, usually occurs through intrinsic or extrinsic apoptotic pathways. The caspases involved in apoptosis can be split into two groups, the initiator caspases such as caspase-9 and effector caspases such as caspase-3. Effector caspases are activated by initiator caspases and mediate many of the morphological cellular changes associated with apoptosis (1).Calcium is an important signaling ion involved in the regulation of many physiological as well as pathological cellular responses (2). In the pancreas, we have shown that Ca2+ signals elicit enzyme secretion (3), apoptosis (46), and pathological intracellular activation of digestive enzymes (7). As such, there must be mechanisms in place by which the cell can differentiate between apoptotic and non-apoptotic Ca2+ signals.The spatiotemporal pattern of calcium signaling is crucial for the specificity of cellular responses. For example, repetitive cytosolic calcium spikes confined to the apical region of the pancreatic acinar cell are elicited by physiological stimulation with acetylcholine (ACh) or cholecystokinin (CCK) and result in physiological secretion of zymogen granules (8, 9). However, a sustained global increase in free cytosolic Ca2+ induced by supramaximal stimulation with CCK, which resembles prolonged hyperstimulation of pancreatic acinar cells in the pathophysiology of acute pancreatitis, can lead to premature activation of digestive enzymes and vacuole formation within the cell (1012). Alternatively, global repetitive calcium spikes induced in the pancreatic acinar cell in response to oxidant stress can lead to induction of the mitochondrial permeability transition pore (mPTP)4 and apoptosis (4, 5, 13).To understand the role of calcium in apoptosis, several investigators have examined the influence of intracellular stores on the molding of calcium signals that lead to cell death (1416). It has been well established in a range of cell types that the endoplasmic reticulum (ER) is the major intracellular calcium store required for induction of apoptosis. Pinton et al. (17) have shown that decreasing ER Ca2+ concentration with tBuBHQ increased HeLa cell survival in response to oxidant stress induced by ceramide. Scorrano and Korsmeyer (18) also observed that double knock-out Bax and Bak (pro-apoptotic proteins) mouse fibroblasts displayed a reduced resting concentration of ER Ca2+ compared with wild type and were resistant to induction of apoptosis by various stimulants, including ceramide. These important studies strongly suggest that the concentration of Ca2+ in the ER is a critical determinant of cellular susceptibility to apoptotic stimuli in the cell types studied.A key event in early apoptosis is permeabilization of the mitochondrial membrane. The mPTP is a pore whose molecular composition is still debated (19). Activation of an open pore state can result in swelling of the mitochondrial matrix and release of the apoptogenic proteins from the intermembrane space (20).One important activator of the mPTP is Ca2+ (2022), a function which implicates Ca2+ in the initiation of apoptosis (23, 24). Once Ca2+ is released from the ER into the cytoplasm, mitochondria take up part of the released Ca2+ to prevent propagation of large calcium waves (2527). This influx is followed by calcium efflux from the mitochondria back into the cytosol (28, 29). An increase in mitochondrial Ca2+ concentration in response to physiological stimuli induces increased activity of the mitochondrial respiratory chain and the synthesis of ATP to meet with increasing energy demands on the cell. When mitochondria are exposed to a pathological overload of calcium, opening of the mPTP is triggered, leading to mitochondrial dysfunction and eventually cell death. The mechanism through which calcium can trigger mPTP opening is still unclear and may involve cyclophilin D (30) and voltage-dependent anion channel (31). The mitochondria are endowed with selective and efficient calcium uptake (a calcium-selective uniporter) and release mechanisms (Ca2+/Na exchanger, Ca2+/H+ exchanger, and mPTP) (16, 29, 32, 33).Oxidant stress is a well known inducer of apoptosis in several cell types (34) and is thought to play an important role in the pathogenesis of acute pancreatitis (35). We have used the quinone compound menadione to induce oxidative stress in the pancreatic acinar cell. Menadione is metabolized by flavoprotein reductase to semiquinone and then is oxidized back to quinone, resulting in generation of superoxide anion radicals, hydrogen peroxide, and other reactive oxygen species (ROS) (36). In vivo, menadione causes depolarization and swelling of the mitochondria (37). In pancreatic acinar cells, treatment with menadione not only produces an increase in ROS, but has also been found to evoke cytosolic Ca2+ responses, mPTP opening, activation of caspases and apoptotic cell death (4, 5). When cells were pretreated with the calcium chelator BAPTA-AM, menadione was unable to induce apoptosis, indicating that oxidant stress-induced apoptosis in the pancreatic acinar cell is highly calcium-dependent. Here we show that in pancreatic acinar cells, oxidative stress-induced apoptosis is strongly dependent on the Ca2+ concentration within mitochondria at the time of ROS production.  相似文献   

15.
The flesh-eating bacterium group A Streptococcus (GAS) binds and activates human plasminogen, promoting invasive disease. Streptococcal surface enolase (SEN), a glycolytic pathway enzyme, is an identified plasminogen receptor of GAS. Here we used mass spectrometry (MS) to confirm that GAS SEN is octameric, thereby validating in silico modeling based on the crystal structure of Streptococcus pneumoniae α-enolase. Site-directed mutagenesis of surface-located lysine residues (SENK252 + 255A, SENK304A, SENK334A, SENK344E, SENK435L, and SENΔ434–435) was used to examine their roles in maintaining structural integrity, enzymatic function, and plasminogen binding. Structural integrity of the GAS SEN octamer was retained for all mutants except SENK344E, as determined by circular dichroism spectroscopy and MS. However, ion mobility MS revealed distinct differences in the stability of several mutant octamers in comparison with wild type. Enzymatic analysis indicated that SENK344E had lost α-enolase activity, which was also reduced in SENK334A and SENΔ434–435. Surface plasmon resonance demonstrated that the capacity to bind human plasminogen was abolished in SENK252 + 255A, SENK435L, and SENΔ434–435. The lysine residues at positions 252, 255, 434, and 435 therefore play a concerted role in plasminogen acquisition. This study demonstrates the ability of combining in silico structural modeling with ion mobility-MS validation for undertaking functional studies on complex protein structures.Streptococcus pyogenes (group A Streptococcus, GAS)8 is a common bacterial pathogen, causing over 700 million human disease episodes each year (1). These range from serious life-threatening invasive diseases including necrotizing fasciitis and streptococcal toxic shock-like syndrome to non-invasive infections like pharyngitis and pyoderma. Invasive disease, in combination with postinfection immune sequelae including rheumatic heart disease and acute poststreptococcal glomerulonephritis, account for over half a million deaths each year (1). Although a resurgence of GAS invasive infections has occurred in western countries since the mid-1980s, disease burden is much greater in developing countries and indigenous populations of developed nations, where GAS infections are endemic (24).GAS is able to bind human plasminogen and activate the captured zymogen to the serine protease plasmin (517). The capacity of GAS to do this plays a critical role in virulence and invasive disease initiation (3, 1719). The plasminogen activation system in humans is an important and highly regulated process that is responsible for breakdown of extracellular matrix components, dissolution of blood clots, and cell migration (20, 21). Plasminogen is a 92-kDa zymogen that circulates in human plasma at a concentration of 2 μm (22). It consists of a binding region of five homologous triple loop kringle domains and an N-terminal serine protease domain that flank the Arg561–Val562 site (23), where it is cleaved by tissue plasminogen activator and urokinase plasminogen activator to yield the active protease plasmin (20, 23). GAS also has the ability to activate human plasminogen by secreting the virulence determinant streptokinase. Streptokinase forms stable complexes with plasminogen or plasmin, both of which exhibit plasmin activity (20, 24). Activation of plasminogen by the plasmin(ogen)-streptokinase complex circumvents regulation by the host plasminogen activation inhibitors, α2-antiplasmin and α2-macroglobulin (11, 20). GAS can bind the plasmin(ogen)-streptokinase complex and/or plasmin(ogen) directly via plasmin(ogen) receptors at the bacterial cell surface (6). These receptors include the plasminogen-binding group A streptococcal M-like protein (PAM) (25), the PAM-related protein (19), glyceraldehyde-3-phosphate dehydrogenase (GAPDH; also known as streptococcal plasmin receptor, Plr, or streptococcal surface dehydrogenase) (9, 26), and streptococcal surface enolase (SEN or α-enolase) (27). Interactions with these GAS receptors occurs via lysine-binding sites within the kringle domains of plasminogen (6).In addition to its ability to bind human plasminogen, SEN is primarily the glycolytic enzyme that converts 2-phosphoglycerate to phosphoenolpyruvate (2729). SEN is abundantly expressed in the cytosol of most bacterial species but has also been identified as a surface-located protein in GAS and other bacteria including pneumococci, despite lacking classical cell surface protein motifs such as a signal sequence, membrane-spanning domain, or cell-wall anchor motif (27, 28, 30, 31). The interaction between SEN and plasminogen is reported to be facilitated by the two C-terminal lysine residues at positions 434 and 435 (27, 32). In contrast, an internal binding motif containing lysines at positions 252 and 255 in the closely related α-enolase of Streptococcus pneumoniae has been shown to play a pivotal role in the acquisition of plasminogen in this bacterial species (33). The octameric pneumococcal α-enolase structure consists of a tetramer of dimers. Hence, potential binding sites could be buried in the interface between subunits. In fact, the crystal structure of S. pneumoniae α-enolase revealed that the two C-terminal lysine residues are significantly less exposed than the internal plasminogen-binding motif (34).In this study, we constructed an in silico model of GAS SEN, based on the pneumococcal octameric α-enolase crystal structure, and validated this model using ion mobility (IM) mass spectrometry (MS). Site-directed mutagenesis followed by structural and functional analyses revealed that Lys344 plays a crucial role in structural integrity and enzymatic function. Furthermore, we demonstrate that the plasminogen-binding motif residues Lys252 and Lys255 and the C-terminal Lys434 and Lys435 residues are located adjacently in the GAS SEN structure and play a concerted role in the binding of human plasminogen.  相似文献   

16.
17.
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   

18.
The double-headed arrowhead protease inhibitors API-A and -B from the tubers of Sagittaria sagittifolia (Linn) feature two distinct reactive sites, unlike other members of their family. Although the two inhibitors have been extensively characterized, the identities of the two P1 residues in both API-A and -B remain controversial. The crystal structure of a ternary complex at 2.48 Å resolution revealed that the two trypsins bind on opposite sides of API-A and are 34 Å apart. The overall fold of API-A belongs to the β-trefoil fold and resembles that of the soybean Kunitz-type trypsin inhibitors. The two P1 residues were unambiguously assigned as Leu87 and Lys145, and their identities were further confirmed by site-directed mutagenesis. Reactive site 1, composed of residues P5 Met83 to P5′ Ala92, adopts a novel conformation with the Leu87 completely embedded in the S1 pocket even though it is an unfavorable P1 residue for trypsin. Reactive site 2, consisting of residues P5 Cys141 to P5′ Glu150, binds trypsin in the classic mode by employing a two-disulfide-bonded loop. Analysis of the two binding interfaces sheds light on atomic details of the inhibitor specificity and also promises potential improvements in enzyme activity by engineering of the reactive sites.Protease inhibitors (PIs)4 are ubiquitously distributed in all organisms, including plants, animals, and microorganisms (1). They play vital roles in regulating their corresponding proteases, which are involved in many biological processes such as protein digestion, cell signal transmission, inflammation, apoptosis, blood coagulation, and embryogenesis (2). The clinical applications of PIs are widespread, and there is great interest in developing more potent therapeutic PIs for treating human diseases related to cancer (3), pancreatitis (4), thrombosis (5), and AIDS (6). To this end, the soybean Kunitz-type serine proteases inhibitors have been extensively studied (1, 711). The inhibitors of this family generally contain 170–200 residues and have two disulfide bonds. Most members have only one reactive site located in the region of residues 60–70 (7, 10, 1214). However, a few members possess two reactive sites that simultaneously bind two protease molecules and are thus termed double-headed inhibitors (1518). All of these inhibitors are classified into family I3 of peptidase inhibitors (19). Most members are further grouped into subfamily I3A. However, the double-headed arrowhead PIs API-A and -B are grouped in subfamily I3B because of their very low sequence similarity to other members (19). In contrast to other double-headed PIs such as the Bowman-Birk and ovomucoid inhibitors, which have two identical reactive sites that have evolved by domain shuffling and gene duplication (1, 2025), both API-A and -B have two distinct reactive sites.API-A and -B were first purified from the tubers of Sagittaria sagittifolia (Linn) in 1979 (26). Both consist of 179 residues with three disulfide bonds and can inhibit a variety of serine proteases, including trypsin, chymotrypsin, and porcine tissue kallikrein (17, 2628). Although the sequence identity of API-A and -B is as high as 91%, their inhibitory specificities differ. The former can bind one molecule of trypsin and one molecule of chymotrypsin, whereas the latter can simultaneously bind two molecules of trypsin (26). The two P1 residues of the reactive sites of API-A and -B were first predicted to be Lys44 and Arg76 based on their surrounding sequences, which are similar to those of the reactive sites of bovine pancreas trypsin inhibitor and soybean Kunitz trypsin inhibitor (29). However, their identities were later revised to Arg76 and Leu87 (for API-A) or Lys87 (for API-B) based on results from sited-directed mutagenesis studies (30).To clarify these controversies, we solved the crystal structure of API-A in complex with two molecules of bovine trypsin. To the best of our knowledge, this is the first report on the three-dimensional structure of the double-headed Kunitz-type trypsin inhibitor in complex with two molecules of protease. On the basis of this structure, the two P1 residues have now been identified as Leu87 and Lys145 for reactive site 1 (RS1) and 2 (RS2), respectively. The results were further confirmed by site-directed mutagenesis. It was earlier shown that the first P1 residue Leu87 interacts preferentially with chymotrypsin (30). In our structure, Leu87 is snugly embedded in the S1 pocket of trypsin, as a consequence of the broad interface contributed by the surrounding residues. Comprehensive analyses of the two reactive site interfaces have provided functional insights into the novel inhibitory patterns of this unique double-headed protease inhibitor.  相似文献   

19.
Although PTIP is implicated in the DNA damage response, through interactions with 53BP1, the function of PTIP in the DNA damage response remain elusive. Here, we show that RNF8 controls DNA damage-induced nuclear foci formation of PTIP, which in turn regulates 53BP1 localization to the DNA damage sites. In addition, SMC1, a substrate of ATM, could not be phosphorylated at the DNA damage sites in the absence of PTIP. The PTIP-dependent pathway is important for DNA double strand breaks repair and DNA damage-induced intra-S phase checkpoint activation. Taken together, these results suggest that the role of PTIP in the DNA damage response is downstream of RNF8 and upstream of 53BP1. Thus, PTIP regulates 53BP1-dependent signaling pathway following DNA damage.The DNA damage response pathways are signal transduction pathways with DNA damage sensors, mediators, and effectors, which are essential for maintaining genomic stability (13). Following DNA double strand breaks, histone H2AX at the DNA damage sites is rapidly phosphorylated by ATM/ATR/DNAPK (410), a family homologous to phosphoinositide 3-kinases (11, 12). Subsequently, phospho-H2AX (γH2AX) provides the platform for accumulation of a larger group of DNA damage response factors, such as MDC1, BRCA1, 53BP1, and the MRE11·RAD50·NBS1 complex (13, 14), at the DNA damage sites. Translocalization of these proteins to the DNA double strand breaks (DSBs)3 facilitates DNA damage checkpoint activation and enhances the efficiency of DNA damage repair (14, 15).Recently, PTIP (Pax2 transactivation domain-interacting protein, or Paxip) has been identified as a DNA damage response protein and is required for cell survival when exposed to ionizing radiation (IR) (1, 1618). PTIP is a 1069-amino acid nuclear protein and has been originally identified in a yeast two-hybrid screening as a partner of Pax2 (19). Genetic deletion of the PTIP gene in mice leads to early embryonic lethality at embryonic day 8.5, suggesting that PTIP is essential for early embryonic development (20). Structurally, PTIP contains six tandem BRCT (BRCA1 carboxyl-terminal) domains (1618, 21). The BRCT domain is a phospho-group binding domain that mediates protein-protein interactions (17, 22, 23). Interestingly, the BRCT domain has been found in a large number of proteins involved in the cellular response to DNA damages, such as BRCA1, MDC1, and 53BP1 (7, 2429). Like other BRCT domain-containing proteins, upon exposure to IR, PTIP forms nuclear foci at the DSBs, which is dependent on its BRCT domains (1618). By protein affinity purification, PTIP has been found in two large complexes. One includes the histone H3K4 methyltransferase ALR and its associated cofactors, the other contains DNA damage response proteins, including 53BP1 and SMC1 (30, 31). Further experiments have revealed that DNA damage enhances the interaction between PTIP and 53BP1 (18, 31).To elucidate the DNA damage response pathways, we have examined the upstream and downstream partners of PTIP. Here, we report that PTIP is downstream of RNF8 and upstream of 53BP1 in response to DNA damage. Moreover, PTIP and 53BP1 are required for the phospho-ATM association with the chromatin, which phosphorylates SMC1 at the DSBs. This PTIP-dependent pathway is involved in DSBs repair.  相似文献   

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