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 437 毫秒
1.
TANK-binding kinase 1 (TBK1) is a key component of Toll-like receptor-dependent and -independent signaling pathways. In response to microbial components, TBK1 activates interferon regulatory factor 3 (IRF3) and cytokine expression. Here we show that TBK1 is a novel target of the γ134.5 protein, a virulence factor whose expression is regulated in a temporal fashion. Remarkably, the γ134.5 protein is required to inhibit IRF3 phosphorylation, nuclear translocation, and the induction of antiviral genes in infected cells. When expressed in mammalian cells, the γ134.5 protein forms complexes with TBK1 and disrupts the interaction of TBK1 and IRF3, which prevents the induction of interferon and interferon-stimulated gene promoters. Down-regulation of TBK1 requires the amino-terminal domain. In addition, unlike wild type virus, a herpes simplex virus mutant lacking γ134.5 replicates efficiently in TBK1-/- cells but not in TBK1+/+ cells. Addition of exogenous interferon restores the antiviral activity in both TBK1-/- and TBK+/+ cells. Hence, control of TBK1-mediated cell signaling by the γ134.5 protein contributes to herpes simplex virus infection. These results reveal that TBK1 plays a pivotal role in limiting replication of a DNA virus.Herpes simplex virus 1 (HSV-1)3 is a large DNA virus that establishes latent or lytic infection, in which the virus triggers innate immune responses. In HSV-infected cells, a number of antiviral mechanisms operate in a cell type- and time-dependent manner (1). In response to double-stranded RNA (dsRNA), Toll-like receptor 3 (TLR3) recruits an adaptor TIR domain-containing adaptor inducing IFN-β and stimulates cytokine expression (2, 3). In the cytoplasm, RNA helicases, RIG-I (retinoid acid-inducible gene-I), and MDA5 (melanoma differentiation associated gene 5) recognize intracellular viral 5′-triphosphate RNA or dsRNA (2, 4). Furthermore, a DNA-dependent activator of IFN-regulatory factor (DAI) senses double-stranded DNA in the cytoplasm and induces cytokine expression (5). There is also evidence that viral entry induces antiviral programs independent of TLR and RIG-I pathways (6). While recognizing distinct viral components, these innate immune pathways relay signals to the two IKK-related kinases, TANK-binding kinase 1 (TBK1) and inducible IκB kinase (IKKi) (2).The IKK-related kinases function as essential components that phosphorylate IRF3 (interferon regulatory factor 3), as well as the closely related IRF7, which translocates to the nucleus and induces antiviral genes, such as interferon-α/β and ISG56 (interferon-stimulated gene 56) (7, 8). TBK1 is constitutively expressed, whereas IKKi is engaged as an inducible gene product of innate immune signaling (9, 10). IRF3 activation is attenuated in TBK1-deficient but not in IKKi-deficient cells (11, 12). Its activation is completely abolished in double-deficient cells (12), suggesting a partially redundant function of TBK1 and IKKi. Indeed, IKKi also negatively regulates the STAT-signaling pathway (13). TBK1/IKKi interacts with several proteins, such as TRAF family member-associated NF-κB activator (TANK), NAP1 (NAK-associated protein 1), similar to NAP1TBK1 adaptor (SINTBAD), DNA-dependent activator of IFN-regulatory factors (DAI), and secretory protein 5 (Sec5) in host cells (5, 1418). These interactions are thought to regulate TBK1/IKKi, which delineates innate as well as adaptive immune responses.Upon viral infection, expression of HSV proteins interferes with the induction of antiviral immunity. When treated with UV or cycloheximide, HSV induces an array of antiviral genes in human lung fibroblasts (19, 20). Furthermore, an HSV mutant, with deletion in immediate early protein ICP0, induces ISG56 expression (21). Accordingly, expression of ICP0 inhibits the induction of antiviral programs mediated by IRF3 or IRF7 (2123). However, although ICP0 negatively regulates IFN-β expression, it is not essential for this effect (24). In HSV-infected human macrophages or dendritic cells, an immediate early protein ICP27 is required to suppress cytokine induction involving IRF3 (25). In this context, it is notable that an HSV mutant, lacking a leaky late gene γ134.5, replicates efficiently in cells devoid of IFN-α/β genes (26). Additionally, the γ134.5 null mutant induces differential cytokine expression as compared with wild type virus (27). Thus, HSV modulation of cytokine expression is a complex process that involves multiple viral components. Currently, the molecular mechanism governing this event is unclear. In this study, we show that HSV γ134.5 targets TBK1 and inhibits antiviral signaling. The data herein reveal a previously unrecognized mechanism by which γ134.5 facilitates HSV replication.  相似文献   

2.
Classic and novel protein kinase C (PKC) isozymes contain two zinc finger motifs, designated “C1a” and “C1b” domains, which constitute the recognition modules for the second messenger diacylglycerol (DAG) or the phorbol esters. However, the individual contributions of these tandem C1 domains to PKC function and, reciprocally, the influence of protein context on their function remain uncertain. In the present study, we prepared PKCδ constructs in which the individual C1a and C1b domains were deleted, swapped, or substituted for one another to explore these issues. As isolated fragments, both the δC1a and δC1b domains potently bound phorbol esters, but the binding of [3H]phorbol 12,13-dibutyrate ([3H]PDBu) by the δC1a domain depended much more on the presence of phosphatidylserine than did that of the δC1b domain. In intact PKCδ, the δC1b domain played the dominant role in [3H]PDBu binding, membrane translocation, and down-regulation. A contribution from the δC1a domain was nonetheless evident, as shown by retention of [3H]PDBu binding at reduced affinity, by increased [3H]PDBu affinity upon expression of a second δC1a domain substituting for the δC1b domain, and by loss of persistent plasma membrane translocation for PKCδ expressing only the δC1b domain, but its contribution was less than predicted from the activity of the isolated domain. Switching the position of the δC1b domain to the normal position of the δC1a domain (or vice versa) had no apparent effect on the response to phorbol esters, suggesting that the specific position of the C1 domain within PKCδ was not the primary determinant of its activity.One of the essential steps for protein kinase C (PKC)2 activation is its translocation from the cytosol to the membranes. For conventional (α, βI, βII, and γ) and novel (δ, ε, η, and θ) PKCs, this translocation is driven by interaction with the lipophilic second messenger sn-1,2-diacylglycerol (DAG), generated from phosphatidylinositol 4,5-bisphosphate upon the activation of receptor-coupled phospholipase C or indirectly from phosphatidylcholine via phospholipase D (1). A pair of zinc finger structures in the regulatory domain of the PKCs, the “C1” domains, are responsible for the recognition of the DAG signal. The DAG-C1 domain-membrane interaction is coupled to a conformational change in PKC, both causing the release of the pseudosubstrate domain from the catalytic site to activate the enzyme and triggering the translocation to the membrane (2). By regulating access to substrates, PKC translocation complements the intrinsic enzymatic specificity of PKC to determine its substrate profile.The C1 domain is a highly conserved cysteine-rich motif (∼50 amino acids), which was first identified in PKC as the interaction site for DAG or phorbol esters (3). It possesses a globular structure with a hydrophilic binding cleft at one end surrounded by hydrophobic residues. Binding of DAG or phorbol esters to the C1 domain caps the hydrophilic cleft and forms a continuous hydrophobic surface favoring the interaction or penetration of the C1 domain into the membrane (4). In addition to the novel and classic PKCs, six other families of proteins have also been identified, some of whose members possess DAG/phorbol ester-responsive C1 domains. These are the protein kinase D (5), the chimaerin (6), the munc-13 (7), the RasGRP (guanyl nucleotide exchange factors for Ras and Rap1) (8), the DAG kinase (9), and the recently characterized MRCK (myotonic dystrophy kinase-related Cdc42-binding kinase) families (10). Of these C1 domain-containing proteins, the PKCs have been studied most extensively and are important therapeutic targets (11). Among the drug candidates in clinical trials that target PKC, a number such as bryostatin 1 and PEP005 are directed at the C1 domains of PKC rather than at its catalytic site.Both the classic and novel PKCs contain in their N-terminal regulatory region tandem C1 domains, C1a and C1b, which bind DAG/phorbol ester (12). Multiple studies have sought to define the respective roles of these two C1 domains in PKC regulation, but the issue remains unclear. Initial in vitro binding measurements with conventional PKCs suggested that 1 mol of phorbol ester bound per mole of PKC (13-15). On the other hand, Stubbs et al., using a fluorescent phorbol ester analog, reported that PKCα bound two ligands per PKC (16). Further, site-directed mutagenesis of the C1a and C1b domains of intact PKCα indicated that the C1a and C1b domains played equivalent roles for membrane translocation in response to phorbol 12-myristate 13-acetate (PMA) and (-)octylindolactam V (17). Likewise, deletion studies indicated that the C1a and C1b domains of PKCγ bound PDBu equally with high potency (3, 18). Using a functional assay with PKCα expression in yeast, Shieh et al. (19) deleted individual C1 domains and reported that C1a and C1b were both functional and equivalent upon stimulation by PMA, with either deletion causing a similar reduction in potency of response, whereas for mezerein the response depended essentially on the C1a domain, with much weaker response if only the C1b domain was present. Using isolated C1 domains, Irie et al. (20) suggested that the C1a domain of PKCα but not those of PKCβ or PKCγ bound [3H]PDBu preferentially; different ligands showed a generally similar pattern but with different extents of selectivity. Using synthesized dimeric bisphorbols, Newton''s group reported (21) that, although both C1 domains of PKCβII are oriented for potential membrane interaction, only one C1 domain bound ligand in a physiological context.In the case of novel PKCs, many studies have been performed on PKCδ to study the equivalency of the twin C1 domains. The P11G point mutation of the C1a domain, which caused a 300-fold loss of binding potency in the isolated domain (22), had little effect on the phorbol ester-dependent translocation of PKCδ in NIH3T3 cells, whereas the same mutation of the C1b caused a 20-fold shift in phorbol ester potency for inducing translocation, suggesting a major role of the C1b domain for phorbol ester binding (23). A secondary role for the C1a domain was suggested, however, because mutation in the C1a domain as well as the C1b domain caused a further 7-fold shift in potency. Using the same mutations in the C1a and C1b domains, Bögi et al. (24) found that the binding selectivity for the C1a and C1b domains of PKCδ appeared to be ligand-dependent. Whereas PMA and the indole alkaloids indolactam and octylindolactam were selectively dependent on the C1b domain, selectivity was not observed for mezerein, the 12-deoxyphorbol 13-monoesters prostratin and 12-deoxyphorbol 13-phenylacetate, and the macrocyclic lactone bryostatin 1 (24). In in vitro studies using isolated C1a and C1b domains of PKCδ, Cho''s group (25) described that the two C1 domains had opposite affinities for DAG and phorbol ester; i.e. the C1a domain showed high affinity for DAG and the C1b domain showed high affinity for phorbol ester. No such difference in selectivity was observed by Irie et al. (20).PKC has emerged as a promising therapeutic target both for cancer and for other conditions, such as diabetic retinopathy or macular degeneration (26-30). Kinase inhibitors represent one promising approach for targeting PKC, and enzastaurin, an inhibitor with moderate selectivity for PKCβ relative to other PKC isoforms (but still with activity on some other non-PKC kinases) is currently in multiple clinical trials. An alternative strategy for drug development has been to target the regulatory C1 domains of PKC. Strong proof of principle for this approach is provided by multiple natural products, e.g. bryostatin 1 and PEP005, which are likewise in clinical trials and which are directed at the C1 domains. A potential advantage of this approach is the lesser number of homologous targets, <30 DAG-sensitive C1 domains compared with over 500 kinases, as well as further opportunities for specificity provided by the diversity of lipid environments, which form a half-site for ligand binding to the C1 domain. Because different PKC isoforms may induce antagonistic activities, inhibition of one isoform may be functionally equivalent to activation of an antagonistic isoform (31).Along with the benzolactams (20, 32), the DAG lactones have provided a powerful synthetic platform for manipulating ligand: C1 domain interactions (31). For example, the DAG lactone derivative 130C037 displayed marked selectivity among the recombinant C1a and C1b domains of PKCα and PKCδ as well as substantial selectivity for RasGRP relative to PKCα (33). Likewise, we have shown that a modified DAG lactone (dioxolanones) can afford an additional point of contact in ligand binding to the C1b domain of PKCδ (34). Such studies provide clear examples that ligand-C1 domain interactions can be manipulated to yield novel patterns of recognition. Further selectivity might be gained with bivalent compounds, exploiting the spacing and individual characteristics of the C1a and C1b domains (35). A better understanding of the differential roles of the two C1 domains in PKC regulation is critical for the rational development of such compounds. In this study, by molecularly manipulating the C1a or C1b domains in intact PKCδ, we find that both the C1a and C1b domains play important roles in PKCδ regulation. The C1b domain is predominant for ligand binding and for membrane translocation of the whole PKCδ molecule. The C1a domain of intact PKCδ plays only a secondary role in ligand binding but stabilizes the PKCδ molecule at the plasma membrane for downstream signaling. In addition, we show that the effect of the individual C1 domains of PKCδ does not critically depend on their position within the regulatory domain.  相似文献   

3.
The paralyzed zebrafish strain relaxed carries a null mutation for the skeletal muscle dihydropyridine receptor (DHPR) β1a subunit. Lack of β1a results in (i) reduced membrane expression of the pore forming DHPR α1S subunit, (ii) elimination of α1S charge movement, and (iii) impediment of arrangement of the DHPRs in groups of four (tetrads) opposing the ryanodine receptor (RyR1), a structural prerequisite for skeletal muscle-type excitation-contraction (EC) coupling. In this study we used relaxed larvae and isolated myotubes as expression systems to discriminate specific functions of β1a from rather general functions of β isoforms. Zebrafish and mammalian β1a subunits quantitatively restored α1S triad targeting and charge movement as well as intracellular Ca2+ release, allowed arrangement of DHPRs in tetrads, and most strikingly recovered a fully motile phenotype in relaxed larvae. Interestingly, the cardiac/neuronal β2a as the phylogenetically closest, and the ancestral housefly βM as the most distant isoform to β1a also completely recovered α1S triad expression and charge movement. However, both revealed drastically impaired intracellular Ca2+ transients and very limited tetrad formation compared with β1a. Consequently, larval motility was either only partially restored (β2a-injected larvae) or not restored at all (βM). Thus, our results indicate that triad expression and facilitation of 1,4-dihydropyridine receptor (DHPR) charge movement are common features of all tested β subunits, whereas the efficient arrangement of DHPRs in tetrads and thus intact DHPR-RyR1 coupling is only promoted by the β1a isoform. Consequently, we postulate a model that presents β1a as an allosteric modifier of α1S conformation enabling skeletal muscle-type EC coupling.Excitation-contraction (EC)3 coupling in skeletal muscle is critically dependent on the close interaction of two distinct Ca2+ channels. Membrane depolarizations of the myotube are sensed by the voltage-dependent 1,4-dihydropyridine receptor (DHPR) in the sarcolemma, leading to a rearrangement of charged amino acids (charge movement) in the transmembrane segments S4 of the pore-forming DHPR α1S subunit (1, 2). This conformational change induces via protein-protein interaction (3, 4) the opening of the sarcoplasmic type-1 ryanodine receptor (RyR1) without need of Ca2+ influx through the DHPR (5). The release of Ca2+ from the sarcoplasmic reticulum via RyR1 consequently induces muscle contraction. The protein-protein interaction mechanism between DHPR and RyR1 requires correct ultrastructural targeting of both channels. In Ca2+ release units (triads and peripheral couplings) of the skeletal muscle, groups of four DHPRs (tetrads) are coupled to every other RyR1 and hence are geometrically arranged following the RyR-specific orthogonal arrays (6).The skeletal muscle DHPR is a heteromultimeric protein complex, composed of the voltage-sensing and pore-forming α1S subunit and auxiliary subunits β1a, α2δ-1, and γ1 (7). While gene knock-out of the DHPR γ1 subunit (8, 9) and small interfering RNA knockdown of the DHPR α2δ-1 subunit (10-12) have indicated that neither subunit is essential for coupling of the DHPR with RyR1, the lack of the α1S or of the intracellular β1a subunit is incompatible with EC coupling and accordingly null model mice die perinatally due to asphyxia (13, 14). β subunits of voltage-gated Ca2+ channels were repeatedly shown to be responsible for the facilitation of α1 membrane insertion and to be potent modulators of α1 current kinetics and voltage dependence (15, 16). Whether the loss of EC coupling in β1-null mice was caused by decreased DHPR membrane expression or by the lack of a putative specific contribution of the β subunit to the skeletal muscle EC coupling apparatus (17, 18) was not clearly resolved. Recently, other β-functions were identified in skeletal muscle using the β1-null mutant zebrafish relaxed (19, 20). Like the β1-knock-out mouse (14) zebrafish relaxed is characterized by complete paralysis of skeletal muscle (21, 22). While β1-knock-out mouse pups die immediately after birth due to respiratory paralysis (14), larvae of relaxed are able to survive for several days because of oxygen and metabolite diffusion via the skin (23). Using highly differentiated myotubes that are easy to isolate from these larvae, the lack of EC coupling could be described by quantitative immunocytochemistry as a moderate ∼50% reduction of α1S membrane expression although α1S charge movement was nearly absent, and, most strikingly, as the complete lack of the arrangement of DHPRs in tetrads (19). Thus, in skeletal muscle the β subunit enables EC coupling by (i) enhancing α1S membrane targeting, (ii) facilitating α1S charge movement, and (iii) enabling the ultrastructural arrangement of DHPRs in tetrads.The question arises, which of these functions are specific for the skeletal muscle β1a and which ones are rather general properties of Ca2+ channel β subunits. Previous reconstitution studies made in the β1-null mouse system (24, 25) using different β subunit constructs (26) did not allow differentiation between β-induced enhancement of non-functional α1S membrane expression and the facilitation of α1S charge movement, due to the lack of information on α1S triad expression levels. Furthermore, the β-induced arrangement of DHPRs in tetrads was not detected as no ultrastructural information was obtained.In the present study, we established zebrafish mutant relaxed as an expression system to test different β subunits for their ability to restore skeletal muscle EC coupling. Using isolated myotubes for in vitro experiments (19, 27) and complete larvae for in vivo expression studies (28-31) and freeze-fracture electron microscopy, a clear differentiation between the major functional roles of β subunits was feasible in the zebrafish system. The cloned zebrafish β1a and a mammalian (rabbit) β1a were shown to completely restore all parameters of EC coupling when expressed in relaxed myotubes and larvae. However, the phylogenetically closest β subunit to β1a, the cardiac/neuronal isoform β2a from rat, as well as the ancestral βM isoform from the housefly (Musca domestica), could recover functional α1S membrane insertion, but led to very restricted tetrad formation when compared with β1a, and thus to impaired DHPR-RyR1 coupling. This impairment caused drastic changes in skeletal muscle function.The present study shows that the enhancement of functional α1S membrane expression is a common function of all the tested β subunits, from β1a to even the most distant βM, whereas the effective formation of tetrads and thus proper skeletal muscle EC coupling is an exclusive function of the skeletal muscle β1a subunit. In context with previous studies, our results suggest a model according to which β1a acts as an allosteric modifier of α1S conformation. Only in the presence of β1a, the α1S subunit is properly folded to allow RyR1 anchoring and thus skeletal muscle-type EC coupling.  相似文献   

4.
The novel α1D L-type Ca2+ channel is expressed in supraventricular tissue and has been implicated in the pacemaker activity of the heart and in atrial fibrillation. We recently demonstrated that PKA activation led to increased α1D Ca2+ channel activity in tsA201 cells by phosphorylation of the channel protein. Here we sought to identify the phosphorylated PKA consensus sites on the α1 subunit of the α1D Ca2+ channel by generating GST fusion proteins of the intracellular loops, N terminus, proximal and distal C termini of the α1 subunit of α1D Ca2+ channel. An in vitro PKA kinase assay was performed for the GST fusion proteins, and their phosphorylation was assessed by Western blotting using either anti-PKA substrate or anti-phosphoserine antibodies. Western blotting showed that the N terminus and C terminus were phosphorylated. Serines 1743 and 1816, two PKA consensus sites, were phosphorylated by PKA and identified by mass spectrometry. Site directed mutagenesis and patch clamp studies revealed that serines 1743 and 1816 were major functional PKA consensus sites. Altogether, biochemical and functional data revealed that serines 1743 and 1816 are major functional PKA consensus sites on the α1 subunit of α1D Ca2+ channel. These novel findings provide new insights into the autonomic regulation of the α1D Ca2+ channel in the heart.L-type Ca2+ channels are essential for the generation of normal cardiac rhythm, for induction of rhythm propagation through the atrioventricular node and for the contraction of the atrial and ventricular muscles (15). L-type Ca2+ channel is a multisubunit complex including α1, β and α2/δ subunits (57). The α1 subunit contains the voltage sensor, the selectivity filter, the ion conduction pore, and the binding sites for all known Ca2+ channel blockers (69). While α1C Ca2+ channel is expressed in the atria and ventricles of the heart (1013), expression of α1D Ca2+ channel is restricted to the sinoatrial (SA)2 and atrioventricular (AV) nodes, as well as in the atria, but not in the adult ventricles (2, 3, 10).Only recently it has been realized that α1D along with α1C Ca2+ channels contribute to L-type Ca2+ current (ICa-L) and they both play important but unique roles in the physiology/pathophysiology of the heart (69). Compared with α1C, α1D L-type Ca2+ channel activates at a more negative voltage range and shows slower current inactivation during depolarization (14, 15). These properties may allow α1D Ca2+ channel to play critical roles in SA and AV nodes function. Indeed, α1D Ca2+ channel knock-out mice exhibit significant SA dysfunction and various degrees of AV block (12, 1619).The modulation of α1C Ca2+ channel by cAMP-dependent PKA phosphorylation has been extensively studied, and the C terminus of α1 was identified as the site of the modulation (2022). Our group was the first to report that 8-bromo-cAMP (8-Br-cAMP), a membrane-permeable cAMP analog, increased α1D Ca2+ channel activity using patch clamp studies (2). However, very little is known about potential PKA phosphorylation consensus motifs on the α1D Ca2+ channel. We therefore hypothesized that the C terminus of the α1 subunit of the α1D Ca2+ channel mediates its modulation by cAMP-dependent PKA pathway.  相似文献   

5.
Cell death can be divided into the anti-inflammatory process of apoptosis and the pro-inflammatory process of necrosis. Necrosis, as apoptosis, is a regulated form of cell death, and Poly-(ADP-Ribose) Polymerase-1 (PARP-1) and Receptor-Interacting Protein (RIP) 1/3 are major mediators. We previously showed that absence or inhibition of PARP-1 protects mice from nephritis, however only the male mice. We therefore hypothesized that there is an inherent difference in the cell death program between the sexes. We show here that in an immune-mediated nephritis model, female mice show increased apoptosis compared to male mice. Treatment of the male mice with estrogens induced apoptosis to levels similar to that in female mice and inhibited necrosis. Although PARP-1 was activated in both male and female mice, PARP-1 inhibition reduced necrosis only in the male mice. We also show that deletion of RIP-3 did not have a sex bias. We demonstrate here that male and female mice are prone to different types of cell death. Our data also suggest that estrogens and PARP-1 are two of the mediators of the sex-bias in cell death. We therefore propose that targeting cell death based on sex will lead to tailored and better treatments for each gender.  相似文献   

6.
Activation of protein kinase C (PKC) promotes the salvage pathway of ceramide formation, and acid sphingomyelinase has been implicated, in part, in providing substrate for this pathway (Zeidan, Y. H., and Hannun, Y. A. (2007) J. Biol. Chem. 282, 11549–11561). In the present study, we examined whether acid β-glucosidase 1 (GBA1), which hydrolyzes glucosylceramide to form lysosomal ceramide, was involved in PKC-regulated formation of ceramide from recycled sphingosine. Glucosylceramide levels declined after treatment of MCF-7 cells with a potent PKC activator, phorbol 12-myristate 13-acetate (PMA). Silencing GBA1 by small interfering RNAs significantly attenuated acid glucocerebrosidase activity and decreased PMA-induced formation of ceramide by 50%. Silencing GBA1 blocked PMA-induced degradation of glucosylceramide and generation of sphingosine, the source for ceramide biosynthesis. Reciprocally, forced expression of GBA1 increased ceramide levels. These observations indicate that GBA1 activation can generate the source (sphingosine) for PMA-induced formation of ceramide through the salvage pathway. Next, the role of PKCδ, a direct effector of PMA, in the formation of ceramide was determined. By attenuating expression of PKCδ, cells failed to trigger PMA-induced alterations in levels of ceramide, sphingomyelin, and glucosylceramide. Thus, PKCδ activation is suggested to stimulate the degradation of both sphingomyelin and glucosylceramide leading to the salvage pathway of ceramide formation. Collectively, GBA1 is identified as a novel source of regulated formation of ceramide, and PKCδ is an upstream regulator of this pathway.Sphingolipids are abundant components of cellular membranes, many of which are emerging as bioactive lipid mediators thought to play crucial roles in cellular responses (1, 2). Ceramide, a central sphingolipid, serves as the main precursor for various sphingolipids, including glycosphingolipids, gangliosides, and sphingomyelin. Regulation of formation of ceramide has been demonstrated through the action of three major pathways: the de novo pathway (3, 4), the sphingomyelinase pathway (5), and the salvage pathway (68). The latter plays an important role in constitutive sphingolipid turnover by salvaging long-chain sphingoid bases (sphingosine and dihydrosphingosine) that serve as sphingolipid backbones for ceramide and dihydroceramide as well as all complex sphingolipids (Fig. 1A).Open in a separate windowFIGURE 1.The scheme of the sphingosine salvage pathway of ceramide formation and inhibition of PMA induction of ceramide by fumonisin B1. A, the scheme of the sphingosine salvage pathway of ceramide formation. B, previously published data as to effects of fumonisin B1 on ceramide mass profiles (23) are re-plotted as a PMA induction of ceramide. In brief, MCF-7 cells were pretreated with or without 100 μm fumonisin B1 for 2 h followed by treatment with 100 nm PMA for 1 h. Lipids were extracted, and then the levels of ceramide species were determined by high-performance liquid chromatography-tandem mass spectrometry. Results are expressed as sum of increased mass of ceramide species. Dotted or open columns represents C16-ceramide or sum of other ceramide species (C14-ceramide, C18-ceramide, C18:1-ceramide, C20-ceramide, C24-ceramide, and C24:1-ceramide), respectively. The data represent mean ± S.E. of three to five values.Metabolically, ceramide is also formed from degradation of glycosphingolipids (Fig. 1A) usually in acidic compartments, the lysosomes and/or late endosomes (9). The stepwise hydrolysis of complex glycosphingolipids eventually results in the formation of glucosylceramide, which in turn is converted to ceramide by the action of acid β-glucosidase 1 (GBA1)2 (9, 10). Severe defects in GBA1 activity cause Gaucher disease, which is associated with aberrant accumulation of the lipid substrates (1014). On the other hand, sphingomyelin is cleaved by acid sphingomyelinase to also form ceramide (15, 16). Either process results in the generation of lysosomal ceramide that can then be deacylated by acid ceramidase (17), releasing sphingosine that may escape the lysosome (18). The released sphingosine may become a substrate for either sphingosine kinases or ceramide synthases, forming sphingosine 1-phosphate or ceramide, respectively (3, 1921).In a related line of investigation, our studies (20, 22, 23) have begun to implicate protein kinase Cs (PKC) as upstream regulators of the sphingoid base salvage pathway resulting in ceramide synthesis. Activation of PKCs by the phorbol ester (PMA) was shown to stimulate the salvage pathway resulting in increases in ceramide. All the induced ceramide was inhibited by pretreatment with a ceramide synthase inhibitor, fumonisin B1, but not by myriocin, thus negating acute activation of the de novo pathway and establishing a role for ceramide synthesis (20, 23). Moreover, labeling studies also implicated the salvage pathway because PMA induced turnover of steady state-labeled sphingolipids but did not affect de novo labeled ceramide in pulse-chase experiments.Moreover, PKCδ, among PKC isoforms, was identified as an upstream molecule for the activation of acid sphingomyelinase in the salvage pathway (22). Interestingly, the PKCδ isoform induced the phosphorylation of acid sphingomyelinase at serine 508, leading to its activation and consequent formation of ceramide. The activation of acid sphingomyelinase appeared to contribute to ∼50% of the salvage pathway-induced increase in ceramide (28) (also, see Fig. 4C). This raised the possibility that distinct routes of ceramide metabolism may account for the remainder of ceramide generation. In this study, we investigated glucocerebrosidase GBA1 as a candidate for one of the other routes accounting for PKC-regulated salvage pathway of ceramide formation.Open in a separate windowFIGURE 4.Effects of knockdown of lysosomal enzymes on the generation of ceramide after PMA treatment. A, MCF-7 cells were transfected with 5 nm siRNAs of each of four individual sequences (SCR, GBA1-a, GBA1-b, and GBA1-c) for 48 h and then stimulated with 100 nm PMA for 1 h. Lipids were extracted, and then the levels of the C16-ceramide species were determined by high-performance liquid chromatography-tandem mass spectrometry. The data represent mean ± S.E. of three to nine values. B, MCF-7 cells were transfected with 5 nm siRNAs of SCR or GBA1-a (GBA1) for 48 h and then stimulated with 100 nm PMA for 1 h. Lipids were extracted, and then the levels of individual ceramide species were determined by high-performance liquid chromatography-tandem mass spectrometry. The data represent mean ± S.E. of three to five values. C14-Cer, C14-ceramide; C16-Cer, C16-ceramide; C18-Cer; C18-ceramide; C18:1-Cer, C18:1-ceramide; C20-Cer, C20-ceramide; C20-Cer, C24-ceramide; C24:1-Cer, C24:1-ceramide. C, MCF-7 cells were transfected with 5 nm siRNAs of SCR, acid sphingomyelinase (ASM), or GBA1-a (GBA1) for 48 h following stimulation with (PMA) or without (Control) 100 nm PMA for 1 h. Lipids were extracted, and then the levels of ceramide species were determined by high-performance liquid chromatography-tandem mass spectrometry. Levels of C16-ceramide are shown. The data represent mean ± S.E. of four to five values. Significant changes from SCR-transfected cells treated with PMA are shown in A–C (*, p < 0.02; **, p < 0.05; ***, p < 0.01).  相似文献   

7.
Excessive accumulation of β-amyloid peptides in the brain is a major cause for the pathogenesis of Alzheimer disease. β-Amyloid is derived from β-amyloid precursor protein (APP) through sequential cleavages by β- and γ-secretases, whose enzymatic activities are tightly controlled by subcellular localization. Delineation of how intracellular trafficking of these secretases and APP is regulated is important for understanding Alzheimer disease pathogenesis. Although APP trafficking is regulated by multiple factors including presenilin 1 (PS1), a major component of the γ-secretase complex, and phospholipase D1 (PLD1), a phospholipid-modifying enzyme, regulation of intracellular trafficking of PS1/γ-secretase and β-secretase is less clear. Here we demonstrate that APP can reciprocally regulate PS1 trafficking; APP deficiency results in faster transport of PS1 from the trans-Golgi network to the cell surface and increased steady state levels of PS1 at the cell surface, which can be reversed by restoring APP levels. Restoration of APP in APP-deficient cells also reduces steady state levels of other γ-secretase components (nicastrin, APH-1, and PEN-2) and the cleavage of Notch by PS1/γ-secretase that is more highly correlated with cell surface levels of PS1 than with APP overexpression levels, supporting the notion that Notch is mainly cleaved at the cell surface. In contrast, intracellular trafficking of β-secretase (BACE1) is not regulated by APP. Moreover, we find that PLD1 also regulates PS1 trafficking and that PLD1 overexpression promotes cell surface accumulation of PS1 in an APP-independent manner. Our results clearly elucidate a physiological function of APP in regulating protein trafficking and suggest that intracellular trafficking of PS1/γ-secretase is regulated by multiple factors, including APP and PLD1.An important pathological hallmark of Alzheimer disease (AD)4 is the formation of senile plaques in the brains of patients. The major components of those plaques are β-amyloid peptides (Aβ), whose accumulation triggers a cascade of neurodegenerative steps ending in formation of senile plaques and intraneuronal fibrillary tangles with subsequent neuronal loss in susceptible brain regions (1, 2). Aβ is proteolytically derived from the β-amyloid precursor protein (APP) through sequential cleavages by β-secretase (BACE1), a novel membrane-bound aspartyl protease (3, 4), and by γ-secretase, a high molecular weight complex consisting of at least four components: presenilin (PS), nicastrin (NCT), anterior pharynx-defective-1 (APH-1), and presenilin enhancer-2 (PEN-2) (5, 6). APP is a type I transmembrane protein belonging to a protein family that includes APP-like protein 1 (APLP1) and 2 (APLP2) in mammals (7, 8). Full-length APP is synthesized in the endoplasmic reticulum (ER) and transported through the Golgi apparatus. Most secreted Aβ peptides are generated within the trans-Golgi network (TGN), also the major site of steady state APP in neurons (911). APP can be transported to the cell surface in TGN-derived secretory vesicles if not proteolyzed to Aβ or an intermediate metabolite. At the cell surface APP is either cleaved by α-secretase to produce soluble sAPPα (12) or reinternalized for endosomal/lysosomal degradation (13, 14). Aβ may also be generated in endosomal/lysosomal compartments (15, 16). In contrast to neurotoxic Aβ peptides, sAPPα possesses neuroprotective potential (17, 18). Thus, the subcellular distribution of APP and proteases that process it directly affect the ratio of sAPPα to Aβ, making delineation of the mechanisms responsible for regulating trafficking of all of these proteins relevant to AD pathogenesis.Presenilin (PS) is a critical component of the γ-secretase. Of the two mammalian PS gene homologues, PS1 and PS2, PS1 encodes the major form (PS1) in active γ-secretase (19, 20). Nascent PSs undergo endoproteolytic cleavage to generate an amino-terminal fragment (NTF) and a carboxyl-terminal fragment (CTF) to form a functional PS heterodimer (21). Based on observations that PSs possess two highly conserved aspartate residues indispensable for γ-secretase activity and that specific transition state analogue γ-secretase inhibitors bind to PS1 NTF/CTF heterodimers (5, 22), PSs are believed to be the catalytic component of the γ-secretase complex. PS assembles with three other components, NCT, APH-1, and PEN-2, to form the functional γ-secretase (5, 6). Strong evidence suggests that PS1/γ-secretase resides principally in the ER, early Golgi, TGN, endocytic and intermediate compartments, most of which (except the TGN) are not major subcellular sites for APP (23, 24). In addition to generating Aβ and cleaving APP to release the APP intracellular domain, PS1/γ-secretase cleaves other substrates such as Notch (25), cadherin (26), ErbB4 (27), and CD44 (28), releasing their respective intracellular domains. Interestingly, PS1/γ-secretase cleavage of different substrates seems to occur at different subcellular compartments; APP is mainly cleaved at the TGN and early endosome domains, whereas Notch is predominantly cleaved at the cell surface (9, 11, 29). Thus, perturbing intracellular trafficking of PS1/γ-secretase may alter interactions between PS1/γ-secretase and APP, contributing to either abnormal Aβ generation and AD pathogenesis or decreased access of PS1/γ-secretase to APP such that Aβ production is reduced. However, mechanisms regulating PS1/γ-secretase trafficking warrant further investigation.In addition to participating in γ-secretase activity, PS1 regulates intracellular trafficking of several membrane proteins, including other γ-secretase components (nicastrin, APH-1, and PEN-2) and the substrate APP (reviewed in Ref. 30). Intracellular APP trafficking is highly regulated and requires other factors such as mint family members and SorLA (2). Moreover, we recently found that phospholipase D1 (PLD1), a phospholipid-modifying enzyme that regulates membrane trafficking events, can interact with PS1, and can regulate budding of APP-containing vesicles from the TGN and delivery of APP to the cell surface (31, 32). Interestingly, Kamal et al. (33) identified an axonal membrane compartment that contains APP, BACE1, and PS1 and showed that fast anterograde axonal transport of this compartment is mediated by APP and kinesin-I, implying a traffic-regulating role for APP. Increased APP expression is also shown to decrease retrograde axonal transport of nerve growth factor (34). However, whether APP indeed regulates intracellular trafficking of proteins including BACE1 and PS1/γ-secretase requires further validation. In the present study we demonstrate that intracellular trafficking of PS1, as well as that of other γ-secretase components, but not BACE1, is regulated by APP. APP deficiency promotes cell surface delivery of PS1/γ-secretase complex and facilitates PS1/γ-secretase-mediated Notch cleavage. In addition, we find that PLD1 also regulates intracellular trafficking of PS1 through a different mechanism and more potently than APP.  相似文献   

8.
9.
Proteolytic processing of amyloid precursor protein (APP) by β- and γ-secretases generates β-amyloid (Aβ) peptides, which accumulate in the brains of individuals affected by Alzheimer disease. Detergent-resistant membrane microdomains (DRM) rich in cholesterol and sphingolipid, termed lipid rafts, have been implicated in Aβ production. Previously, we and others reported that the four integral subunits of the γ-secretase associate with DRM. In this study we investigated the mechanisms underlying DRM association of γ-secretase subunits. We report that in cultured cells and in brain the γ-secretase subunits nicastrin and APH-1 undergo S-palmitoylation, the post-translational covalent attachment of the long chain fatty acid palmitate common in lipid raft-associated proteins. By mutagenesis we show that nicastrin is S-palmitoylated at Cys689, and APH-1 is S-palmitoylated at Cys182 and Cys245. S-Palmitoylation-defective nicastrin and APH-1 form stable γ-secretase complexes when expressed in knock-out fibroblasts lacking wild type subunits, suggesting that S-palmitoylation is not essential for γ-secretase assembly. Nevertheless, fractionation studies show that S-palmitoylation contributes to DRM association of nicastrin and APH-1. Moreover, pulse-chase analyses reveal that S-palmitoylation is important for nascent polypeptide stability of both proteins. Co-expression of S-palmitoylation-deficient nicastrin and APH-1 in cultured cells neither affects Aβ40, Aβ42, and AICD production, nor intramembrane processing of Notch and N-cadherin. Our findings suggest that S-palmitoylation plays a role in stability and raft localization of nicastrin and APH-1, but does not directly modulate γ-secretase processing of APP and other substrates.Alzheimer disease is the most common among neurodegenerative diseases that cause dementia. This debilitating disorder is pathologically characterized by the cerebral deposition of 39–42 amino acid peptides termed Aβ, which are generated by proteolytic processing of amyloid precursor protein (APP)2 by β- and γ-secretases (1, 2). The β-site APP cleavage enzyme 1 cleaves full-length APP within its luminal domain to generate a secreted ectodomain leaving behind a C-terminal fragment (β-CTF). γ-Secretase cleaves β-CTF within the transmembrane domain to release Aβ and APP intracellular C-terminal domain (AICD). γ-Secretase is a multiprotein complex, comprising at least four subunits: presenilins (PS1 and PS2), nicastrin, APH-1, and PEN-2 for its activity (3). PS1 is synthesized as a 42–43-kDa polypeptide and undergoes highly regulated endoproteolytic processing within the large cytoplasmic loop domain connecting putative transmembrane segments 6 and 7 to generate stable N-terminal (NTF) and C-terminal fragments (CTF) by an uncharacterized proteolytic activity (4). This endoproteolytic event has been identified as the activation step in the process of PS1 maturation as it assembles with other γ-secretase subunits (3). Nicastrin is a heavily glycosylated type I membrane protein with a large ectodomain that has been proposed to function in substrate recognition and binding (5), but this putative function has not been confirmed by others (6). APH-1 is a seven-transmembrane protein encoded by two human or three rodent genes that are alternatively spliced (7). Although PS1 (or PS2), nicastrin, APH-1, and PEN-2 are sufficient for γ-secretase processing of APP, a type I membrane protein, termed p23 (also referred toTMP21), was recently identified as a γ-secretase component that modulates γ-secretase activity and regulates secretory trafficking of APP (8, 9).A growing number of type I integral membrane proteins has been identified as γ-secretase substrates within the last few years, including Notch1 homologues, Notch ligands, Delta and Jagged, cell adhesion receptors N- and E-cadherins, low density lipoprotein receptor-related protein, ErbB-4, netrin receptor DCC, and others (10). Mounting evidence suggests that APP processing occurs within cholesterol- and sphingolipid-enriched lipid rafts, which are biochemically defined as detergentresistant membrane microdomains (DRM) (11, 12). Previously we reported that each of the γ-secretase subunits localizes in lipid rafts in post-Golgi and endosome membranes enriched in syntaxin 6 (13). Moreover, loss of γ-secretase activity by gene deletion or exposure to γ-secretase inhibitors results in the accumulation of APP CTFs in lipid rafts indicating that cleavage of APP CTFs likely occurs in raft microdomains (14). In contrast, CTFs derived from Notch1, Jagged2, N-cadherin, and DCC are processed by γ-secretase in non-raft membranes (14). The mechanisms underlying association of γ-secretase subunits with lipid rafts need further clarification to elucidate spatial segregation of amyloidogenic processing of APP in membrane microdomains.Post-translational S-palmitoylation is increasingly recognized as a potential mechanism for regulating raft association, stability, intracellular trafficking, and function of several cytosolic and transmembrane proteins (1517). S-palmitoylation refers to the addition of 16-carbon palmitoyl moiety to certain cysteine residues through thioester linkage. Cysteines close to transmembrane domains or membrane-associated domains in non-integral membrane proteins are preferred S-palmitoylation sites, although no conserved motif has been identified (18). Palmitoylation modifies numerous neuronal proteins, including postsynaptic density protein PSD-95 (19), a-amino-3-hydroxyl-5-methyl-4-isoxazole propionic acid receptors (20), nicotinic α7 receptors (21), neuronal t-SNAREs SNAP-25, synaptobrevin 2 and synaptogagmin (22, 23), neuronal growth-associated protein GAP-43 (24), protein kinase CLICK-III (CL3)/CaMKIγ (25), β-secretase (26), and Huntingtin (27). Although palmitoylation can occur in vitro without the involvement of an enzyme, a family of palmitoyltransferases that specifically catalyze S-palmitoylation has been identified (28, 29).In this study, we have identified S-palmitoylation of γ-secretase subunits nicastrin and APH-1, and characterized its role on DRM association, protein stability, and γ-secretase enzyme activities. We show that nicastrin is S-palmitoylated at Cys689, and APH-1 at Cys182 and Cys245. Mutagenesis of palmitoylation sites results in increased degradation of nascent nicastrin and APH-1 polypeptides and reduced association with DRM. Nevertheless, in cultured cells overexpression of S-palmitoylation-deficient nicastrin and APH-1 does not modulate γ-secretase processing of APP or other substrates.  相似文献   

10.
Accumulation of the amyloid β (Aβ) peptide derived from the proteolytic processing of amyloid precursor protein (APP) is the defining pathological hallmark of Alzheimer disease. We previously demonstrated that the C-terminal 37 amino acids of lipoprotein receptor-related protein (LRP) robustly promoted Aβ generation independent of FE65 and specifically interacted with Ran-binding protein 9 (RanBP9). In this study we found that RanBP9 strongly increased BACE1 cleavage of APP and Aβ generation. This pro-amyloidogenic activity of RanBP9 did not depend on the KPI domain or the Swedish APP mutation. In cells expressing wild type APP, RanBP9 reduced cell surface APP and accelerated APP internalization, consistent with enhanced β-secretase processing in the endocytic pathway. The N-terminal half of RanBP9 containing SPRY-LisH domains not only interacted with LRP but also with APP and BACE1. Overexpression of RanBP9 resulted in the enhancement of APP interactions with LRP and BACE1 and increased lipid raft association of APP. Importantly, knockdown of endogenous RanBP9 significantly reduced Aβ generation in Chinese hamster ovary cells and in primary neurons, demonstrating its physiological role in BACE1 cleavage of APP. These findings not only implicate RanBP9 as a novel and potent regulator of APP processing but also as a potential therapeutic target for Alzheimer disease.The major defining pathological hallmark of Alzheimer disease (AD)2 is the accumulation of amyloid β protein (Aβ), a neurotoxic peptide derived from β- and γ-secretase cleavages of the amyloid precursor protein (APP). The vast majority of APP is constitutively cleaved in the middle of the Aβ sequence by α-secretase (ADAM10/TACE/ADAM17) in the non-amyloidogenic pathway, thereby abrogating the generation of an intact Aβ peptide. Alternatively, a small proportion of APP is cleaved in the amyloidogenic pathway, leading to the secretion of Aβ peptides (37–42 amino acids) via two proteolytic enzymes, β- and γ-secretase, known as BACE1 and presenilin, respectively (1).The proteolytic processing of APP to generate Aβ requires the trafficking of APP such that APP and BACE1 are brought together in close proximity for β-secretase cleavage to occur. We and others have shown that the low density lipoprotein receptor-related protein (LRP), a multifunctional endocytosis receptor (2), binds to APP and alters its trafficking to promote Aβ generation. The loss of LRP substantially reduces Aβ release, a phenotype that is reversed when full-length (LRP-FL) or truncated LRP is transfected in LRP-deficient cells (3, 4). Specifically, LRP-CT lacking the extracellular ligand binding regions but containing the transmembrane domain and the cytoplasmic tail is capable of rescuing amyloidogenic processing of APP and Aβ release in LRP deficient cells (3). Moreover, the LRP soluble tail (LRP-ST) lacking the transmembrane domain and only containing the cytoplasmic tail of LRP is sufficient to enhance Aβ secretion (5). This activity of LRP-ST is achieved by promoting APP/BACE1 interaction (6), although the precise mechanism is unknown. Although we had hypothesized that one or more NPXY domains in LRP-ST might underlie the pro-amyloidogenic processing of APP, we recently found that the 37 C-terminal residues of LRP (LRP-C37) lacking the NPXY motif was sufficient to robustly promote Aβ production independent of FE65 (7). Because LRP-C37 likely acts by recruiting other proteins, we used the LRP-C37 region as bait in a yeast two-hybrid screen, resulting in the identification of 4 new LRP-binding proteins (7). Among these, we focused on Ran-binding protein 9 (RanBP9) in this study, which we found to play a critical role in the trafficking and processing of APP. RanBP9, also known as RanBPM, acts as a multi-modular scaffolding protein, bridging interactions between the cytoplasmic domains of a variety of membrane receptors and intracellular signaling targets. These include Axl and Sky (8), MET receptor protein-tyrosine kinase (9), and β2-integrin LFA-1 (10). Similarly, RanBP9 interacts with Plexin-A receptors to strongly inhibit axonal outgrowth (11) and functions to regulate cell morphology and adhesion (12, 13). Here we show that RanBP9 robustly promotes BACE1 processing of APP and Aβ generation.  相似文献   

11.
12.
13.
14.
Aggregatibacter actinomycetemcomitans is a gram-negative opportunistic oral pathogen. It is frequently associated with subgingival biofilms of both chronic and aggressive periodontitis, and the diseased sites of the periodontium exhibit increased levels of the proinflammatory mediator interleukin (IL)-1β. Some bacterial species can alter their physiological properties as a result of sensing IL-1β. We have recently shown that this cytokine localizes to the cytoplasm of A. actinomycetemcomitans in co-cultures with organotypic gingival mucosa. However, current knowledge about the mechanism underlying bacterial IL-1β sensing is still limited. In this study, we characterized the interaction of A. actinomycetemcomitans total membrane protein with IL-1β through electrophoretic mobility shift assays. The interacting protein, which we have designated bacterial interleukin receptor I (BilRI), was identified through mass spectrometry and was found to be Pasteurellaceae specific. Based on the results obtained using protein function prediction tools, this protein localizes to the outer membrane and contains a typical lipoprotein signal sequence. All six tested biofilm cultures of clinical A. actinomycetemcomitans strains expressed the protein according to phage display-derived antibody detection. Moreover, proteinase K treatment of whole A. actinomycetemcomitans cells eliminated BilRI forms that were outer membrane specific, as determined through immunoblotting. The protein was overexpressed in Escherichia coli in both the outer membrane-associated form and a soluble cytoplasmic form. When assessed using flow cytometry, the BilRI-overexpressing E. coli cells were observed to bind 2.5 times more biotinylated-IL-1β than the control cells, as detected with avidin-FITC. Overexpression of BilRI did not cause binding of a biotinylated negative control protein. In a microplate assay, soluble BilRI bound to IL-1β, but this binding was not specific, as a control protein for IL-1β also interacted with BilRI. Our findings suggest that A. actinomycetemcomitans expresses an IL-1β-binding surface-exposed lipoprotein that may be part of the bacterial IL-1β-sensing system.  相似文献   

15.
16.
N-Glycosylation of integrin α5β1 plays a crucial role in cell spreading, cell migration, ligand binding, and dimer formation, but the detailed mechanisms by which N-glycosylation mediates these functions remain unclear. In a previous study, we showed that three potential N-glycosylation sites (α5S3–5) on the β-propeller of the α5 subunit are essential to the functional expression of the subunit. In particular, site 5 (α5S5) is the most important for its expression on the cell surface. In this study, the function of the N-glycans on the integrin β1 subunit was investigated using sequential site-directed mutagenesis to remove the combined putative N-glycosylation sites. Removal of the N-glycosylation sites on the I-like domain of the β1 subunit (i.e. the Δ4-6 mutant) decreased both the level of expression and heterodimeric formation, resulting in inhibition of cell spreading. Interestingly, cell spreading was observed only when the β1 subunit possessed these three N-glycosylation sites (i.e. the S4-6 mutant). Furthermore, the S4-6 mutant could form heterodimers with either α5S3-5 or α5S5 mutant of the α5 subunit. Taken together, the results of the present study reveal for the first time that N-glycosylation of the I-like domain of the β1 subunit is essential to both the heterodimer formation and biological function of the subunit. Moreover, because the α5S3-5/β1S4-6 mutant represents the minimal N-glycosylation required for functional expression of the β1 subunit, it might also be useful for the study of molecular structures.Integrin is a heterodimeric glycoprotein that consists of both an α and a β subunit (1). The interaction between integrin and the extracellular matrix is essential to both physiologic and pathologic events, such as cell migration, development, cell viability, immune homeostasis, and tumorigenesis (2, 3). Among the integrin superfamily, β1 integrin can combine with 12 distinct α subunits (α1–11, αv) to form heterodimers, thereby acquiring a wide variety of ligand specificity (1, 4). Integrins are thought to be regulated by inside-out signaling mechanisms that provoke conformational changes, which modulate the affinity of integrin for the ligand (5). However, an increasing body of evidence suggests that cell-surface carbohydrates mediate a variety of interactions between integrin and its extracellular environment, thereby affecting integrin activity and possibly tumor metastasis as well (68).Guo et al. (9) reported that an increase in β1–6-GlcNAc sugar chains on the integrin β1 subunit stimulated cell migration. In addition, elevated sialylation of the β1 subunit, because of Ras-induced STGal-I transferase activity, also induced cell migration (10, 11). Conversely, cell migration and spreading were reduced by the addition of a bisecting GlcNAc, which is a product of N-acetylglucosaminyltransferase III (GnT-III),2 to the α5β1 and α3β1 integrins (12, 13). Alterations of N-glycans on integrins might also regulate their cis interactions with membrane-associated proteins, including the epidermal growth factor receptor, the galectin family, and the tetraspanin family of proteins (1419).In addition to the positive and negative regulatory effects of N-glycan, several research groups have reported that N-glycans must be present on integrin α5β1 for the αβ heterodimer formation and proper integrin-matrix interactions. Consistent with this hypothesis, in the presence of the glycosylation inhibitor, tunicamycin, normal integrin-substrate binding and transport to the cell surface are inhibited (20). Moreover, treatment of purified integrin with N-glycosidase F blocked both the inherent association of the subunits and the interaction between integrin and fibronectin (FN) (21). These results suggest that N-glycosylation is essential to the functional expression of α5β1. However, because integrin α5β1 contains 26 potential N-linked glycosylation sites, 14 in the α subunit and 12 in the β subunit, identification of the sites that are essential to its biological functions is key to understanding the molecular mechanisms by which N-glycans alter integrin function. Recently, our group determined that N-glycosylation of the β-propeller domain on the α5 subunit is essential to both heterodimerization and biological functions of the subunit. Furthermore, we determined that sites 3–5 are the most important sites for α5 subunit-mediated cell spreading and migration on FN (22). The purpose of this study was to clarify the roles of N-glycosylation of the β1 subunit. Therefore, we performed combined substitutions in the putative N-glycosylation sites by replacement of asparagine residues with glutamine residues. We subsequently introduced these mutated genes into β1-deficient epithelial cells (GE11). The results of these mutation experiments revealed that the N-glycosylation sites on the I-like domain of the β1 subunit, sites number 4–6 (S4-6), are essential to both heterodimer formation and biological functions, such as cell spreading.  相似文献   

17.
We have previously shown that residues Tyr-253 and Glu-255 in the serpin antithrombin function as exosites to promote the inhibition of factor Xa and factor IXa when the serpin is conformationally activated by heparin. Here we show that functional exosites can be engineered at homologous positions in a P1 Arg variant of the serpin α1-proteinase inhibitor (α1PI) that does not require heparin for activation. The combined effect of the two exosites increased the association rate constant for the reactions of α1PI with factors Xa and IXa 11–14-fold, comparable with their rate-enhancing effects on the reactions of heparin-activated antithrombin with these proteases. The effects of the engineered exosites were specific, α1PI inhibitor reactions with trypsin and thrombin being unaffected. Mutation of Arg-150 in factor Xa, which interacts with the exosite residues in heparin-activated antithrombin, abrogated the ability of the engineered exosites in α1PI to promote factor Xa inhibition. Binding studies showed that the exosites enhance the Michaelis complex interaction of α1PI with S195A factor Xa as they do with the heparin-activated antithrombin interaction. Replacement of the P4-P2 AIP reactive loop residues in the α1PI exosite variant with a preferred IEG substrate sequence for factor Xa modestly enhanced the reactivity of the exosite mutant inhibitor with factor Xa by ∼2-fold but greatly increased the selectivity of α1PI for inhibiting factor Xa over thrombin by ∼1000-fold. Together, these results show that a specific and selective inhibitor of factor Xa can be engineered by incorporating factor Xa exosite and reactive site recognition determinants in a serpin.The ubiquitous proteins of the serpin superfamily share a common structure and mostly function as inhibitors of intracellular and extracellular serine and cysteine-type proteases in a vast array of physiologic processes (1, 2). Serpins inhibit their target proteases by a suicide substrate inhibition mechanism in which an exposed reactive loop of the serpin is initially recognized as a substrate by the protease. Subsequent cleavage of the reactive loop by the protease up to the acyl-intermediate stage of proteolysis triggers a massive conformational change in the serpin that kinetically traps the acyl-intermediate (3, 4). Although it is well established that serpins recognize their cognate proteases through a specific reactive loop “bait” sequence, it has more recently become clear that serpin exosites outside the reactive loop provide crucial determinants of protease specificity (57). In the case of the blood clotting regulator antithrombin and its target proteases, physiological rates of protease inhibition are only possible with the aid of exosites generated upon activation of the serpin by heparin binding (5). Mutagenesis studies have shown that the antithrombin exosites responsible for promoting the interaction of heparin-activated antithrombin with factor Xa and factor IXa map to two key residues, Tyr-253 and Glu-255, in strand 3 of β-sheet C (8, 9). Parallel mutagenesis studies of factor Xa and factor IXa have shown that the protease residues that interact with the antithrombin exosites reside in the autolysis loop, arginine 150 in this loop being most important (10, 11). The crystal structures of the Michaelis complexes of heparin-activated antithrombin with catalytically inactive S195A variants of thrombin and factor Xa have confirmed that these complexes are stabilized by exosites in antithrombin and in heparin (1214). In particular, the Michaelis complex with S195A factor Xa revealed that Tyr-253 of antithrombin and Arg-150 of factor Xa comprise a critical protein-protein interaction of the antithrombin exosite, in agreement with mutagenesis studies. Binding studies of antithrombin interactions with S195A proteases have shown that the exosites in heparin-activated antithrombin increase the binding affinity for proteases minimally by ∼1000-fold in the Michaelis complex (15, 16).In this study, we have grafted the two exosites in strand 3 of β-sheet C of antithrombin onto their homologous positions in a P1 Arg variant of α1-proteinase inhibitor (α1PI)2 and shown that the exosites are functional in promoting α1PI inhibition of factor Xa and factor IXa. The exosites specifically promote factor Xa and factor IXa inhibition and do not affect the inhibition of trypsin or thrombin. Moreover, mutation of the complementary exosite residue in factor Xa, Arg-150, largely abrogates the rate-enhancing effect of the engineered exosites in α1PI on factor Xa inhibition. Binding studies show that the exosites function by promoting the binding of α1PI and factor Xa in the Michaelis complex. Replacing the P4-P2 residues of the P1 Arg α1PI with an IEG factor Xa recognition sequence modestly enhances the reactivity of the exosite mutant of α1PI with factor Xa and greatly increases the selectivity of the mutant α1PI for inhibiting factor Xa over thrombin. These findings demonstrate that a potent and selective inhibitor of factor Xa can be engineered by grafting exosite and reactive site determinants for the protease on a serpin scaffold.  相似文献   

18.
The success of Mycobacterium tuberculosis (Mtb) as a pathogen rests upon its ability to grow intracellularly in macrophages. Interferon-gamma (IFN-γ) is critical in host defense against Mtb and stimulates macrophage clearance of Mtb through an autophagy pathway. Here we show that the host protein ubiquilin 1 (UBQLN1) promotes IFN-γ-mediated autophagic clearance of Mtb. Ubiquilin family members have previously been shown to recognize proteins that aggregate in neurodegenerative disorders. We find that UBQLN1 can interact with Mtb surface proteins and associates with the bacilli in vitro. In IFN-γ activated macrophages, UBQLN1 co-localizes with Mtb and promotes the anti-mycobacterial activity of IFN-γ. The association of UBQLN1 with Mtb depends upon the secreted bacterial protein, EsxA, which is involved in permeabilizing host phagosomes. In autophagy-deficient macrophages, UBQLN1 accumulates around Mtb, consistent with the idea that it marks bacilli that traffic through the autophagy pathway. Moreover, UBQLN1 promotes ubiquitin, p62, and LC3 accumulation around Mtb, acting independently of the E3 ligase parkin. In summary, we propose a model in which UBQLN1 recognizes Mtb and in turn recruits the autophagy machinery thereby promoting intracellular control of Mtb. Thus, polymorphisms in ubiquilins, which are known to influence susceptibility to neurodegenerative illnesses, might also play a role in host defense against Mtb.  相似文献   

19.
γ-Secretase is known to play a pivotal role in the pathogenesis of Alzheimer disease through production of amyloidogenic Aβ42 peptides. Early onset familial Alzheimer disease mutations in presenilin (PS), the catalytic core of γ-secretase, invariably increase the Aβ42:Aβ40 ratio. However, the mechanism by which these mutations affect γ-secretase complex formation and cleavage specificity is poorly understood. We show that our in vitro assay system recapitulates the effect of PS1 mutations on the Aβ42:Aβ40 ratio observed in cell and animal models. We have developed a series of small molecule affinity probes that allow us to characterize active γ-secretase complexes. Furthermore we reveal that the equilibrium of PS1- and PS2-containing active complexes is dynamic and altered by overexpression of Pen2 or PS1 mutants and that formation of PS2 complexes is positively correlated with increased Aβ42:Aβ40 ratios. These data suggest that perturbations to γ-secretase complex equilibrium can have a profound effect on enzyme activity and that increased PS2 complexes along with mutated PS1 complexes contribute to an increased Aβ42:Aβ40 ratio.β-Amyloid (Aβ)5 peptides are believed to play a causative role in Alzheimer disease (AD). Aβ peptides are generated from the processing of the amyloid precursor protein (APP) by two proteases, β-secretase and γ-secretase. Although γ-secretase generates heterogenous Aβ peptides ranging from 37 to 46 amino acids in length, significant work has focused mainly on the Aβ40 and Aβ42 peptides that are the major constituents of amyloid plaques. γ-Secretase is a multisubunit membrane aspartyl protease comprised of at least four known subunits: presenilin (PS), nicastrin (Nct), anterior pharynx-defective (Aph), and presenilin enhancer 2 (Pen2). Presenilin is thought to contain the catalytic core of the complex (14), whereas Aph and Nct play critical roles in the assembly, trafficking, and stability of γ-secretase as well as substrate recognition (5, 6). Lastly Pen2 facilitates the endoproteolysis of PS into its N-terminal (NTF) and C-terminal (CTF) fragments thereby yielding a catalytically competent enzyme (5, 710). All four proteins (PS, Nct, Aph1, and Pen2) are obligatory for γ-secretase activity in cell and animal models (11, 12). There are two homologs of PS, PS1 and PS2, and three isoforms of Aph1, Aph1aS, Aph1aL, and Aph1b. At least six active γ-secretase complexes have been reported (two presenilins × three Aph1s) (13, 14). The sum of apparent molecular masses of the four proteins (PS1-NTF/CTF ≈ 53 kDa, Nct ≈ 120 kDa, Aph1 ≈ 30 kDa, and Pen2 ≈ 10kDa) is ∼200 kDa. However, active γ-secretase complexes of varying sizes, ranging from 250 to 2000 kDa, have been reported (1519). Recently a study suggested that the γ-secretase complex contains only one of each subunit (20). Collectively these studies suggest that a four-protein complex around 200–250 kDa may be the minimal functional γ-secretase unit with additional cofactors and/or varying stoichiometry of subunits existing in the high molecular weight γ-secretase complexes. CD147 and TMP21 have been found to be associated with the γ-secretase complex (21, 22); however, their role in the regulation of γ-secretase has been controversial (23, 24).Mutations of PS1 or PS2 are associated with familial early onset AD (FAD), although it is debatable whether these familial PS mutations act as “gain or loss of function” alterations in regard to γ-secretase activity (2527). Regardless the overall outcome of these mutations is an increased ratio of Aβ42:Aβ40. Clearly these mutations differentially affect γ-secretase activity for the production of Aβ40 and Aβ42. Despite intensive studies of Aβ peptides and γ-secretase, the molecular mechanism controlling the specificity of γ-secretase activity for Aβ40 and Aβ42 production has not been resolved. It has been found that PS1 mutations affect the formation of γ-secretase complexes (28). However, the precise mechanism by which individual subunits alter the dynamics of γ-secretase complex formation and activity is largely unresolved. A better mechanistic understanding of γ-secretase activity associated with FAD mutations has been hindered by the lack of suitable assays and probes that are necessary to recapitulate the effect of these mutations seen in cell models and to characterize the active γ-secretase complex.In our present studies, we have determined the overall effect of Pen2 and PS1 expression on the dynamics of PS1- and PS2-containing complexes and their association with γ-secretase activity. Using newly developed biotinylated small molecular probes and activity assays, we revealed that expression of Pen2 or PS1 FAD mutants markedly shifts the equilibrium of PS1-containing active complexes to that of PS2-containing complexes and results in an overall increase in the Aβ42:Aβ40 ratio in both stable cell lines and animal models. Our studies indicate that perturbations to the equilibrium of active γ-secretase complexes by an individual subunit can greatly affect the activity of the enzyme. Moreover they serve as further evidence that there are multiple and distinct γ-secretase complexes that can exist within the same cells and that their equilibrium is dynamic. Additionally the affinity probes developed here will facilitate further study of the expression and composition of endogenous active γ-secretase from a variety of model systems.  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