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.
Cell surface receptors of the integrin family are pivotal to cell adhesion and migration. The activation state of heterodimeric αβ integrins is correlated to the association state of the single-pass α and β transmembrane domains. The association of integrin αIIbβ3 transmembrane domains, resulting in an inactive receptor, is characterized by the asymmetric arrangement of a straight (αIIb) and tilted (β3) helix relative to the membrane in congruence to the dissociated structures. This allows for a continuous association interface centered on helix-helix glycine-packing and an unusual αIIb(GFF) structural motif that packs the conserved Phe-Phe residues against the β3 transmembrane helix, enabling αIIb(D723)β3(R995) electrostatic interactions. The transmembrane complex is further stabilized by the inactive ectodomain, thereby coupling its association state to the ectodomain conformation. In combination with recently determined structures of an inactive integrin ectodomain and an activating talin/β complex that overlap with the αβ transmembrane complex, a comprehensive picture of integrin bi-directional transmembrane signaling has emerged.Key words: cell adhesion, membrane protein, integrin, platelet, transmembrane complex, transmembrane signalingThe communication of biological signals across the plasma membrane is fundamental to cellular function. The ubiquitous family of integrin adhesion receptors exhibits the unusual ability to convey signals bi-directionally (outside-in and inside-out signaling), thereby controlling cell adhesion, migration and differentiation.15 Integrins are Type I heterodimeric receptors that consist of large extracellular domains (>700 residues), single-pass transmembrane (TM) domains, and mostly short cytosolic tails (<70 residues). The activation state of heterodimeric integrins is correlated to the association state of the TM domains of their α and β subunits.610 TM dissociation initiated from the outside results in the transmittal of a signal into the cell, whereas dissociation originating on the inside results in activation of the integrin to bind ligands such as extracellular matrix proteins. The elucidation of the role of the TM domains in integrin-mediated adhesion and signaling has been the subject of extensive research efforts, perhaps commencing with the demonstration that the highly conserved GFFKR sequence motif of α subunits (Fig. 1), which closely follows the first charged residue on the intracellular face, αIIb(K989), constrains the receptor to a default low affinity state.11 Despite these efforts, an understanding of this sequence motif had not been reached until such time as the structure of the αIIb TM segment was determined.12 In combination with the structure of the β3 TM segment13 and available mutagenesis data,6,9,10,14,15 this has allowed the first correct prediction of the overall association of an integrin αβ TM complex.12 The predicted association was subsequently confirmed by the αIIbβ3 complex structure determined in phospholipid bicelles,16 as well as by the report of a similar structure based on molecular modeling using disulfide-based structural constraints.17 In addition to the structures of the dissociated and associated αβ TM domains, their membrane embedding was defined12,13,16,18,19 and it was experimentally recognized that, in the context of the native receptor, the TM complex is stabilized by the inactive, resting ectodomain.16 These advances in integrin membrane structural biology are complemented by the recent structures of a resting integrin ectodomain and an activating talin/β cytosolic tail complex that overlap with the αβ TM complex,20,21 allowing detailed insight into integrin bi-directional TM signaling.Open in a separate windowFigure 1Amino acid sequence of integrin αIIb and β3 transmembrane segments and flanking regions. Membrane-embedded residues12,13,16,18,19 are enclosed by a gray box. Residues 991–995 constitute the highly conserved GFFKR sequence motif of integrin α subunits.  相似文献   

2.
3.
4.
Coxsackievirus B3 (CVB3) is a small RNA virus associated with diseases such as myocarditis, meningitis, and pancreatitis. We have previously demonstrated that proteasome inhibition reduces CVB3 replication and attenuates virus-induced myocarditis. However, the underlying mechanisms by which the ubiquitin/proteasome system regulates CVB replication remain unclear. In this study, we investigated the role of REGγ, a member of the 11S proteasome activator, in CVB3 replication. We showed that overexpression of REGγ promoted CVB3 replication but that knockdown of REGγ led to reduced CVB3 replication. We further demonstrated that REGγ-mediated p53 proteolysis contributes, as least in part, to the proviral function of REGγ. Although total protein levels of REGγ remained unaltered after CVB3 infection, virus infection induced a redistribution of REGγ from the nucleus to the cytoplasm, rendering an opportunity for a direct interaction of REGγ with viral proteins and/or host proteins (e.g., p53), which controls viral growth and thereby enhances viral infectivity. Further analyses suggested a potential modification of REGγ by SUMO following CVB3 infection, which was verified by both in vitro and in vivo sumoylation assays. Sumoylation of REGγ may play a role in its nuclear export during CVB3 infection. Taken together, our results present the first evidence that the host REGγ pathway is utilized and modified during CVB3 infection to promote efficient viral replication.Viruses often adapt to the existing host cellular machinery to complete their own life cycle. The ubiquitin/proteasome system (UPS), a primary intracellular protein degradation system in eukaryotic cells, has emerged as a key modulator in viral infectivity and virus-mediated pathogenesis (6).Coxsackievirus B3 (CVB3) is a small RNA virus associated with diseases such as myocarditis, meningitis, and pancreatitis (36). We have previously studied the function and regulation of the UPS in CVB3 infection and CVB3-induced myocarditis (7, 16, 17, 33). We demonstrated that CVB3 utilizes and manipulates the host UPS to achieve successful replication (17, 33). We provided evidence that proteasome inhibition reduces CVB3 replication and attenuates virus-induced myocarditis (7). However, we recognize the potential toxicity of general inhibition of proteasome function as a therapeutic means. Further investigation to identify specific targets within the UPS utilized during CVB3 infection is urgently needed and will allow for more-precise targeting in drug therapy.The 20S proteasome is a multisubunit protease complex responsible for the degradation of misfolded proteins or short-lived regulatory proteins (16, 18). In the absence of proteasome activators, the 20S proteasome is latent and the protein substrates are barred from entering the 20S proteasome (16, 18). There are at least two families of proteasome activators, the 19S proteasome (also known as PA700) and the 11S proteasome (also known as REG or PA28) (16, 18). The 19S activator binds to proteasome to form the 26S proteasome, which primarily performs degradation of proteins in a ubiquitin-dependent manner.The REG activator binds to and activates the proteasome in an ATP-independent manner to promote mainly ubiquitin-independent protein degradation. Three classes of REG have been identified, REGα, REGβ, and REGγ. REGα/β forms a heteroheptamer which is mainly localized to the cytosol (16, 18). The level of REGα/β is inducible by gamma interferon, and the main function of REGα/β has been implicated in major histocompatibility complex (MHC) class I antigen presentation (16, 18). REGγ exists in a homoheptamer and is primarily found in the nucleus (16, 18). Although the functional significance of REGγ has not been fully defined, studies of REGγ-deficient mice reveal a role for REGγ in the regulation of cell cycle progression and cell survival/apoptosis (1, 27). These effects appear to be related to REGγ-mediated degradation of several important intracellular proteins, such as cyclin-dependent kinase inhibitors p21, p16, and p19 (2, 14) and tumor suppressor p53 (43). Moreover, an interaction between the REGγ system and the viral proteins has recently been reported. It was shown that REGγ binds to and regulates the stability and nuclear retention of hepatitis C core protein (26), contributing to hepatitis C core protein-induced insulin resistance and hepatocarcinoma (24, 25).We have previously reported that gene silencing of ubiquitin reduces viral protein synthesis and viral titers (33). However, such inhibitions are not as potent as by proteasome inhibition, suggesting that 11S proteasome-mediated proteasomal degradation may also play a role. In the present study, we seek to further understand the underlying mechanisms by which the UPS regulates CVB3 replication by investigating the interplay between REGγ and CVB3 infection and exploring the potential mechanisms of how REGγ controls CVB3 replication. Here, we provided the first evidence that the host REGγ pathway was utilized and modulated during CVB3 infection to promote efficient viral replication.  相似文献   

5.
6.
7.
8.
9.
Cell migration during wound healing is a complex process that involves the expression of a number of growth factors and cytokines. One of these factors, transforming growth factor-beta (TGFβ) controls many aspects of normal and pathological cell behavior. It induces migration of keratinocytes in wounded skin and of epithelial cells in damaged cornea. Furthermore, this TGFβ-induced cell migration is correlated with the production of components of the extracellular matrix (ECM) proteins and expression of integrins and matrix metalloproteinases (MMPs). MMP digests ECMs and integrins during cell migration, but the mechanisms regulating their expression and the consequences of their induction remain unclear. It has been suggested that MMP-14 activates cellular signaling processes involved in the expression of MMPs and other molecules associated with cell migration. Because of the manifold effects of MMP-14, it is important to understand the roles of MMP-14 not only the cleavage of ECM but also in the activation of signaling pathways.Key words: wound healing, migration, matrix metalloproteinase, transforming growth factor, skin, corneaWound healing is a well-ordered but complex process involving many cellular activities including inflammation, growth factor or cytokine secretion, cell migration and proliferation. Migration of skin keratinocytes and corneal epithelial cells requires the coordinated expression of various growth factors such as platelet-derived growth factor (PDGF), fibroblast growth factor (FGF), transforming growth factor (TGF), keratinocyte growth factor (KGF), hepatocyte growth factor (HGF), insulin-like growth factor (IGF), epidermal growth factor (EGF), small GTPases, and macrophage stimulating protein (reviewed in refs. 1 and 2). The epithelial cells in turn regulate the expression of matrix metalloproteinases (MMPs), extracellular matrix (ECM) proteins and integrins during cell migration.1,3,4 TGF-β is a well-known cytokine involved in processes such as cell growth inhibition, embryogenesis, morphogenesis, tumorigenesis, differentiation, wound healing, senescence and apoptosis (reviewed in refs. 5 and 6). It is also one of the most important cytokines responsible for promoting the migration of skin keratinocytes and corneal epithelial cells.3,6,7TGFβ has two quite different effects on skin keratinocytes: it suppresses their multiplication and promotes their migration. The TGFβ-induced cell growth inhibition is usually mediated by Smad signaling, which upregulates expression of the cell cycle inhibitor p21WAF1/Cip1 or p12CDK2-AP1 in HaCaT skin keratinocyte cells and human primary foreskin keratinocytes.8,9 Keratinocyte migration in wounded skin is associated with strong expression of TGFβ and MMPs,1 and TGFβ stimulates the migration of manually scratched wounded HaCaT cells.10 TGFβ also induces cell migration and inhibits proliferation of injured corneal epithelial cells, whereas it stimulates proliferation of normal corneal epithelial cells via effects on the MAPK family and Smad signaling.2,7 Indeed, skin keratinocytes and corneal epithelial cells display the same two physiological responses to TGFβ during wound healing; cell migration and growth inhibition. However as mentioned above, TGFβ has a different effect on normal cells. For example, it induces the epithelial to mesenchymal transition (EMT) of normal mammary cells and lens epithelial cells.11,12 It also promotes the differentiation of corneal epithelial cells, and induces the fibrosis of various tissues.2,6The MMPs are a family of structurally related zinc-dependent endopeptidases that are secreted into the extracellular environment.13 Members of the MMP family have been classified into gelatinases, stromelysins, collagenases and membrane type-MMPs (MT-MMPs) depending on their substrate specificity and structural properties. Like TGFβ, MMPs influence normal physiological processes including wound healing, tissue remodeling, angiogenesis and embryonic development, as well as pathological conditions such as rheumatoid arthritis, atherosclerosis and tumor invasion.13,14The expression patterns of MMPs during skin and cornea wound healing are well studied. In rats, MMP-2, -3, -9, -11, -13 and -14 are expressed,15 and in mice, MMP-1, -2, -3, -9, -10 and -14 are expressed during skin wound healing.1 MMP-1, -3, -7 and -12 are increased in corneal epithelial cells during Wnt 7a-induced rat cornea wound healing.16 Wound repair after excimer laser keratectomy is characterized by increased expression of MMP-1, -2, -3 and -9 in the rabbit cornea, and MMP-2, -9 in the rat cornea.17,18 The expression of MMP-2 and -9 during skin keratinocyte and corneal epithelial cell migration has been the most thoroughly investigated, and it has been shown that their expression generally depends on the activity of MMP-14. MMP-14 (MT1-MMP) is constitutively anchored to the cell membrane; it activates other MMPs such as MMP-2, and also cleaves various types of ECM molecules including collagens, laminins, fibronectin as well as its ligands, the integrins.13 The latent forms of some cytokines are also cleaved and activated by MMP-14.19 Overexpression of MMP-14 protein was found to stimulate HT1080 human fibrosarcoma cell migration.20 In contrast, the attenuation of MMP-14 expression using siRNA method decreased fibroblast invasiveness,21 angiogenesis of human microvascular endothelial cells,22 and human skin keratinocyte migration.10 The latter effect was shown to result from lowering MMP-9 expression. Other studies have shown that EGF has a critical role in MMP-9 expression during keratinocyte tumorigenesis and migration.23,24 On the other hand, TGFβ modulates MMP-9 production through the Ras/MAPK pathway in transformed mouse keratinocytes and NFκB induces cell migration by binding to the MMP-9 promoter in human skin primary cultures.25,26 Enhanced levels of pro-MMP-9 and active MMP-9 have also been noted in scratched corneal epithelia of diabetic rats.27There is evidence that MMP-14 activates a number of intracellular signaling pathways including the MAPK family pathway, focal adhesion kinase (FAK), Src family, Rac and CD44, during cell migration and tumor invasion.19,20,28 In COS-7 cells, ERK activation is stimulated by overexpression of MMP-14 and is essential for cell migration.29 These observations all indicate that MMP-14 plays an important role in cell migration, not only by regulating the activity or expression of downstream MMPs but also by processing and activating migration-associated molecules such as integrins, ECMs and a variety of intracellular signaling pathays.30Cell migration during wound healing is a remarkably complex phenomenon. TGFβ is just one small component of the overall process of wound healing and yet it triggers a multitude of reactions needed for cell migration. It is important to know what kinds of molecules are expressed when cell migration is initiated, but it is equally important to investigate the roles of these molecules and how their expression is regulated. Despite the availability of some information about how MMPs and signaling molecules can influence each other, much remains to be discovered in this area. It will be especially important to clarify how MMP-14 influences other signaling pathways since its role in cell migration is not restricted to digesting ECM molecules but also includes direct or indirect activation of cellular signaling pathways.  相似文献   

10.
Amyloid beta (Aβ), the putative causative agent in Alzheimer disease, is known to affect glutamate receptor trafficking. Previous studies have shown that Aβ downregulates the surface expression of N-methyl D-aspartate type glutamate receptors (NMDARs) by the activation of STriatal-Enriched protein tyrosine Phosphatase 61 (STEP61). More recent findings confirm that STEP61 plays an important role in Aβ-induced NMDAR endocytosis. STEP levels are elevated in human AD prefrontal cortex and in the cortex of several AD mouse models. The increase in STEP61 levels and activity contribute to the removal of GluN1/GluN2B receptor complexes from the neuronal surface membranes. The elevation of STEP61 is due to disruption in the normal degradation of STEP61 by the ubiquitin proteasome system. Here, we briefly discuss additional studies in support of our hypothesis that STEP61 contributes to aspects of the pathophysiology in Alzheimer''s disease. Exogenous application of Aβ-enriched conditioned medium (7PA2-CM) to wild-type cortical cultures results in a loss of GluN1/GluN2B subunits from neuronal membranes. Abeta-mediated NMDAR internalization does not occur in STEP knock-out cultures, but is rescued by the addition of active TAT-STEP to the cultures prior to Aβ treatment.Key words: Alzheimer disease, amyloid beta, NMDA receptor, protein tyrosine phosphatases, STEP, synaptic plasticityIn Alzheimer disease (AD), the abnormal accumulation of soluble Aβ peptides has a profound impact on cognitive function.1 Aβ peptides disrupt synaptic plasticity, a molecular mechanism involved in learning and memory.2,3 N-methyl D-aspartate type glutamate receptors (NMDAR) play an important role in the development of synaptic strengthening. Aβ downregulates the surface expression of NMDARs by activation of STriatal-Enriched protein tyrosine Phosphatase 61 (STEP61).4 STEP61 is a brain-specific phosphatase that opposes the development of synaptic strengthening.5 STEP61 is present in postsynaptic terminals, immunoprecipitates with the NMDAR complex and decreases NMDA channel function.6,7 The reduced channel function is mediated, at least in part, by an increased internalization of the NMDAR complex, as STEP dephosphorylates the GluN2B subunit at a regulatory tyrosine (tyr1472) leading to NMDAR endocytosis. Knocking down STEP with interfering RNA increases NMDAR trafficking to synaptic membranes.4,8 A previous study suggested that Aβ leads to the activation of STEP through a calcineurin-mediated pathway, which subsequently increased internalization of surface NMDAR.4 A recent study has demonstrated that STEP is also regulated by the ubiquitin proteasome system, and an Aβ-mediated disruption of the proteasome leads to increased STEP61 levels in human Alzheimer''s disease (AD) brains and AD mouse models.9 Taken together, these studies suggest that an increase in the activity of STEP61 contributes to the cognitive deficits in AD by increasing the internalization of NMDAR from synaptic membrane surfaces.  相似文献   

11.
Approved proteasome inhibitors have advanced the treatment of multiple myeloma but are associated with serious toxicities, poor pharmacokinetics, and most with the inconvenience of intravenous administration. We therefore sought to identify novel orally bioavailable proteasome inhibitors with a continuous daily dosing schedule and improved therapeutic window using a unique drug discovery platform. We employed a fluorine-based medicinal chemistry technology to synthesize 14 novel analogs of epoxyketone-based proteasome inhibitors and screened them for their stability, ability to inhibit the chymotrypsin-like proteasome, and antimyeloma activity in vitro. The tolerability, pharmacokinetics, pharmacodynamic activity, and antimyeloma efficacy of our lead candidate were examined in NOD/SCID mice. We identified a tripeptide epoxyketone, FV-162, as a metabolically stable, potent proteasome inhibitor cytotoxic to human myeloma cell lines and primary myeloma cells. FV-162 had limited toxicity and was well tolerated on a continuous daily dosing schedule. Compared with the benchmark oral irreversible proteasome inhibitor, ONX-0192, FV-162 had a lower peak plasma concentration and longer half-life, resulting in a larger area under the curve (AUC). Oral FV-162 treatment induced rapid, irreversible inhibition of chymotrypsin-like proteasome activity in murine red blood cells and inhibited tumor growth in a myeloma xenograft model. Our data suggest that oral FV-162 with continuous daily dosing schedule displays a favorable safety, efficacy, and pharmacokinetic profile in vivo, identifying it as a promising lead for clinical evaluation in myeloma therapy.The ubiquitin–proteasome system is responsible for the regulation and degradation of the majority of the intracellular proteins in eukaryotic cells.1 The 26S proteasome is a multi-subunit protein complex that mediates the proteolytic degradation and turnover of damaged, misfolded or excess proteins that have been polyubiquitylated in the cytoplasm and nucleus.1, 2 The 26S proteasome consists of a 20S core particle, capped by 19S regulatory particles.3, 4 The 19S particle participates in the recognition, processing, unfolding, and translocation of ubiquitylated protein substrates into the 20S core.5 Substrates are then degraded inside the chamber of the barrel-like 20S core particle, where the active sites of multiple β1, β2, and β5 subunits catalyze caspase-like (C-L), trypsin-like (T-L), or chymotrypsin-like (CT-L) proteolysis, respectively.6, 7 Inhibition of the 26S proteasome activity leads to disruption of the cell cycle and induction of apoptosis.8Cancer cells have an increased dependency on the integrity of the ubiquitin–proteasome system machinery compared with normal cells in preclinical studies. This finding is predominantly evident in hematological malignancies, identifying the 26S proteasome as a promising anticancer therapeutic target.9, 10, 11, 12 In particular, cells derived from multiple myeloma are notably sensitive to proteasome inhibition, at least in part, owing to their characteristically high rates of immunoglobulin protein biosynthesis and increased proteasome activity.13, 14, 15 The continuous activity of the proteasome in myeloma cells makes them particularly susceptible to prolonged inhibition.16Bortezomib, the first proteasome inhibitor approved for clinical use, is a dipeptide boronic acid that reversibly binds to the active site of the β5 and β1 subunit to competitively inhibit proteasome function.9, 10, 17 By inhibiting the proteasome, bortezomib acts through multiple cellular pathways that ultimately result in cell cycle arrest and apoptosis.18 Bortezomib is currently approved for the treatment of newly diagnosed, relapsed or refractory multiple myeloma and mantle cell lymphoma.18 Carfilzomib was subsequently developed as a second-generation inhibitor that belongs to the epoxyketone class and irreversibly binds to the active site of the β5 subunit of the proteasome. Carfilzomib is structurally and mechanistically distinct from bortezomib and overcomes bortezomib resistance in multiple myeloma cell lines and in primary multiple myeloma cells from patients.17, 19 Carfilzomib is currently also approved for relapsed and refractory multiple myeloma. ONX-0912 (also known as oprozomib) is another epoxyketone class oral proteasome inhibitor that is an analog of carfilzomib.20, 21 Similar to carfilzomib, ONX-0912 promotes cell death in primary myeloma cells from patients who relapsed after treatment with bortezomib.20 ONX-0912 has advanced into phase I/II trials in hematological and solid malignancies.22, 23, 24Despite their clinical efficacy, treatment with proteasome inhibitors is associated with a number of toxicities, including neuropathy, thrombocytopenia, and cardiotoxicity.25, 26, 27 The toxicity of currently available proteasome inhibitors necessitates administering the drugs in intermittent dosing schedules, typically biweekly. Although intermittent dosing permits proteasome activity in normal tissues during dose holidays, it has been shown to be sub-optimal for therapy in malignant cells.16 Moreover, infrequent administration at relatively high exposures may give rise to undesirable and potentially unnecessary toxicities in normal cells. Potentially, by moderating exposures, an optimized oral proteasome inhibitor with continuous daily dosing could be developed that exploits the high proteasome dependency in malignant cells while sparing normal cells.In the present study, we report the development of FV-162, a novel, metabolically stable and orally bioavailable epoxyketone-based proteasome inhibitor. FV-162 displays potent anticancer activity and maintains a wide differential activity between malignant and normal cells despite a continuous daily dosing schedule in multiple myeloma cell lines, primary patient cells, and animal models. Overall, our results show that FV-162 inhibits the proteasome, displays metabolic stability, and has a favorable toxicity profile.  相似文献   

12.
Jenny M. Woof 《MABS-AUSTIN》2012,4(3):291-293
Fc receptors and their interaction with antibodies will be a major theme at the forthcoming FASEB Science Research Conference on Immunoreceptors to be held in Snowmass this July (details available at www.faseb.org/src/home.aspx, follow the tabs for Immunoreceptors). Since its inception in the mid 1980s, this meeting series has maintained a focus on Fc receptors, and this year’s meeting will be no exception.From a therapeutic viewpoint, there is much to be gained from a detailed understanding of the biology of effector molecules such as Fc receptors and complement. Indeed, knowledge of the interaction of IgG with such molecules has been central to the development of improved mAbs with altered functions and transformed half-lives, tailored for particular therapeutic applications. Examples include mAbs designed to maximise complement recruitment1 or to enhance Fc receptor engagement and triggering of ADCC,2-5 or conversely, variants engineered to be unable to engage complement6 or Fc receptors.7 Glycoengineering of IgG Fc offers an alternative means to modify effector function capabilities,8 while development of IgG mutants that display extended or altered serum half-lives has been driven through exhaustive analysis of the interaction with FcRn.9,10Despite the appreciable advances that have been made in unravelling the various facets of Fc receptor biology, new information pertinent to mAb engineering continues to emerge. A flavour of some of these new advances will be given below. They span novel receptors and receptor roles, structure-function relationships, the molecular architecture of signaling complexes, the influence of the membrane lipid environment and scaffolding interactions, isotype considerations, through to technical innovations likely to inform the field.Remarkably, new receptors that have previously eluded characterization are now being described. These include the IgM receptor, which evidence indicates is a molecule also known as TOPO/Fas apoptotic inhibitory molecule 3 whose gene lies close to other known immunoglobulin receptors on chromosome 1,11 and a receptor for IgD recently documented on basophils.12 Moreover, we are seeing an appreciation of new roles for existing Fc receptors. An example is the demonstration in a transgenic study that human FcγRIIa can trigger active and passive anaphylaxis and airway inflammation. Moreover, human mast cells, monocytes and neutrophils were shown to produce anaphylactogenic mediators when FcγRIIA was engaged.13 Hence IgG may contribute to allergic and anaphylactic reactions in humans by engaging FcγRIIa.Exciting new structural information on Fc receptors and their ligands is emerging. An important example is the solving of the X-ray crystal structure for human FcγRI.14 While the structural information supports a ligand binding mode similar to those of FcγRII or FcγRIII, the FG-loop in domain 2 of FcγRI with its conserved one-residue deletion appears critical for high affinity IgG binding. A second example concerns the high responder/low responder (HR/LR) polymorphisms of FcγRIIa, which are linked to susceptibility to infections, autoimmune diseases, and the efficacy of therapeutic Abs. New insights into these differences have been provided by the recent solving of the structure for the complex of the HR allele with IgG Fc.15 Third, understanding of the human IgE-FcεRI interaction has moved forward significantly through the solving of the X-ray crystal structure of the complex of FcεRI and the entire Fc region of IgE (comprising domains Cε2, Cε3 and Cε4).16 In a final example, the structural basis for the improved efficacy of nonfucosylated mAbs has been investigated.17 The X-ray crystal structure of the complex between nonfucosylated IgG Fc and a soluble form of FcγRIIIa carrying two N-linked glycans showed that one of two receptor glycans interacts with nonfucosylated Fc to stabilize the complex. It is proposed that when the Fc glycan is fucosylated this interaction is inhibited due to steric hindrance and, together with the negative effects of Fc fucosylation on the dynamics of the receptor binding site, this provides a rationale for the improved ADCC displayed by nonfucosylated IgG.A question of interest is precisely how Fc receptors bound to antibody ligands organize themselves within signaling complexes in the cell membrane. Some intriguing clues to this conundrum of molecular architecture are now surfacing. In mast cells, FcεRI molecules loaded with IgE form a synapse when presented with antigen that is mobile within a lipid bilayer, via coalescence into large cholesterol-rich clusters.18 Of particular relevance to the therapeutic setting, clustering of receptors into immune synapses is also seen with FcγR. For instance, during in vivo ADCC mediated by tumor-specific mAb, clustering of FcγR, actin and phosphotyrosines has been noted at contact zones between tumor cells and macrophages or neutrophils.19 The theme of the influence of the membrane lipid domain environment on Fc receptor function is taken up elsewhere. It has been shown, for example, that serine phosphorylation of FcγRI influences membrane mobility and function. The cytoplasmic tail of FcγRI interacts with protein 4.1G,20 and it is proposed that this is mediated via a phosphoserine-dependent mechanism critical for localization of the receptor to lipid rafts.21 With regard to FcγRIIa, a major role for lipid rafts in the regulation of IgG binding to FcγRIIa has been revealed.22 Notably, exclusion of FcγRIIa from lipid raft membrane microdomains is able to suppress IgG binding in myeloid cells.Increased knowledge of the capabilities of Fc receptors specific for other antibody classes is opening up new options for therapy. For example, IgA antibodies may offer a highly useful and efficacious alternative approach of particular relevance to treatment at mucosal sites. Human IgA mAbs have been demonstrated to mediate efficient tumor cell killing23,24 and to have the capability to control certain infectious diseases.25,26 The detailed understanding of functional sites in IgA that has resulted from numerous mutagenesis studies,27 coupled with improved ways to produce and isolate recombinant IgA mAbs28 should facilitate developments toward therapeutics based on this immunoglobulin class. Similarly, recent studies indicate that IgE may serve as an alternative to the classic IgG backbone for therapeutic antibodies.29Finally, technical innovations seem poised to further inform the field and advances are arriving or may be anticipated from techniques such as solution nuclear magnetic resonance (NMR) spectroscopy,30 cryo-electron tomography,31 single particle tracking,32 and ultrasensitive force techniques such as adhesion frequency assays.33,34Interest in Fc receptors continues unabated, and the contribution that the field can make to mAb development and optimisation is unquestionable. The FASEB SRC on Immunoreceptors will serve as a forum for discourse on the above issues and much more, providing invaluable information and networking opportunities for all those interested in ways to maximise the efficacy of mAbs and mAb-based reagents. Registration is open until 24 June 2012.  相似文献   

13.
14.
Protein kinase C (PKC)-ε, a component of the serine/threo-nine PKC family, has been shown to influence the survival and differentiation pathways of normal hematopoietic cells. Here, we have modulated the activity of PKC-ε with specific small molecule activator or inhibitor peptides. PKC-ε inhibitor and activator peptides showed modest effects on HL-60 maturation when added alone, but PKC-ε activator peptide significantly counteracted the pro-maturative activity of tumor necrosis factor (TNF)-α towards the monocytic/macrophagic lineage, as evaluated in terms of CD14 surface expression and morphological analyses. Moreover, while PKC-ε inhibitor peptide showed a reproducible increase of TNF-related apoptosis inducing ligand (TRAIL)-induced apoptosis, PKC-ε activator peptide potently counteracted the pro-apoptotic activity of TRAIL. Taken together, the anti-maturative and anti-apoptotic activities of PKC-ε envision a potentially important proleukemic role of this PKC family member.Key words: acute myeloid leukemia, surface antigens, HL-60 cells, apoptosis, maturation.Activation of all protein kinase C (PKC) family of serine and threonine isoenzymes is associated with binding to the negatively charged phospholipids, phosphatidylserine, while different PKC isozymes have varying sensitivities to Ca2+ and lipid-derived second messengers such as diacylglycerol (Gonelli et al., 2009). Upon activation, PKC isozymes translocate from the soluble to the particulate cell fraction, including cell membrane, nucleus and mitochondria (Gonelli et al., 2009). PKC primary sequence can be broadly separated into two domains: the N-terminal regulatory domain and the conserved C-terminal catalytic domain.The regulatory domain of PKC is composed of the C1 and C2 domains that mediate PKC interactions with second messengers, phospholipids, as well as inter and intramolecular protein-protein interactions. Differences in the order and number of copies of signaling domains, as well as sequence differences that affect binding affinities, result in the distinct activity of each PKC isozyme (Gonelli et al., 2009).In recent years, a series of peptides derived from PKC have been shown to modulate its activity by interfering with critical protein-protein interactions within PKC and between PKC and PKC-binding proteins (Brandman et al., 2007, Souroujon and Mochly-Rosen, 1998). Focusing on PKC-ε isozyme and using a rational approach, one C2-derived peptide that acts as an isozyme-selective activator (Dorn et al., 1999) and another that acts as a selective inhibitor (Johnson et al., 1996) of PKC-ε, have been identified.These findings are particularly interesting since besides being involved in the physiology of normal cardiac (Braun and Mochly-Rosen, 2003, Johnson et al., 1996, Li et al., 2006), hematopoietic (Gobbi et al., 2009, Mirandola et al., 2006, Racke et al., 2001), and neuronal (Borgatti et al., 1996) cell models, mounting experimental evidences have linked altered PKC-ε functions to solid tumor development (Okhrimenko et al., 2005, Gillespie et al., 2005, Lu et al., 2006). Therefore, taking advantage of the recent availability of small molecule peptides able to activate or inhibit specifically PKC-ε by disrupting protein/protein interactions (Dorn et al., 1999, Johnson et al., 1996), which open important therapeutic perspectives, we have investigated the effects of both PKC-ε activator and PKC-ε inhibitor peptides on the maturation and survival of leukemic cells, using as a model system the HL-60 myeloblastic leukemia cell line, which can be induced to undergo terminal differentiation or apoptotic cell death by a variety of chemical and biological agents (Breitman et al., 1980, Zauli et al., 1996).  相似文献   

15.
Eukaryotic DNA polymerase η (Polη) confers ultraviolet (UV) resistance by catalyzing translesion synthesis (TLS) past UV photoproducts. Polη has been studied extensively in budding yeast and mammalian cells, where its interaction with monoubiquitylated proliferating cell nuclear antigen (PCNA) is necessary for its biological activity. Recently, in collaboration with other investigators, our laboratory demonstrated that Arabidopsis thaliana Polη is required for UV resistance in plants. Furthermore, the purified enzyme can perform TLS opposite a cyclobutane pyrimidine dimer and interacts with PCNA. Intriguingly, the biological activity of Polη in a heterologous yeast assay depends on co-expression with Arabidopsis PCNA2 and Polη sequences implicated in binding PCNA or ubiquitin. We suggest that interaction of Arabidopsis Polη with ubiquitylated PCNA2 is required for TLS past UV photoproducts by Polη.Key words: polymerase η, proliferating cell nuclear antigen, translesion synthesis, ubiquitin, Arabidopsis thaliana, ultraviolet radiationUltraviolet (UV)-induced pyrimidine dimers can block the progression of DNA replication forks potentially disrupting the replication machinery and resulting in cell death. For this reason, cells have evolved non-essential, low fidelity DNA polymerases (Pols) capable of copying damaged templates,1,2 a process termed translesion DNA synthesis (TLS). In budding yeast, TLS past UV photoproducts is catalyzed by Polη and Polζ (composed of the Rev3 catalytic and Rev7 accessory subunits), but also involves the Rev1 protein in an as yet undetermined role linked to Polζ.1,3,4 Yeast and human Polη replicates cyclobutane pyrimidine dimers (CPDs), in particular thymine-thymine (TT) CPDs, in a relatively error-free manner whereas Polζ is essential for UV mutagenesis implicating it in error-prone TLS.1,4,5Both UV resistance due to TLS and the polymerases responsible have been well-studied in yeast and mammalian cells over the past decade. Only more recently has evidence emerged that TLS may also contribute to UV resistance in plants. Arabidopsis thaliana POLH, REV1, REV3 and REV7 encode homologs of Polη, Rev1, Rev3 and Rev7, respectively.610 T-DNA insertions in POLH, REV1 or REV3 sensitise root growth to acute UV doses,68,10 and these mutations, as well as inactivation of REV7, increase the sensitivity of whole plants to longer term UV treatment.6,8 Interestingly, polh rev3 double mutants show an additive increase in UV sensitivity over that observed for polh and rev3 single mutants,6,10 potentially pointing to differences in the UV photoproducts bypassed by the two polymerases. That the enhanced UV sensitivity of the mutants may reflect a TLS deficiency is suggested by the finding that purified Arabidopsis Polη catalyzes primer extension and TLS past a TT CPD in vitro.6For TLS to occur, Polη must gain access to the replication machinery arrested at a UV photoproduct. It does so in yeast and mammalian cells by interacting with proliferating cell nuclear antigen (PCNA), the eukaryotic sliding clamp required for processive DNA replication.1,3,11, DNA damage or stalling of the replicative polymerase triggers monoubiquitylation of PCNA at lysine 164 by a complex of the E2 ubiquitin conjugase Rad6 and the E3 ubiquitin ligase Rad18.1,3,11,12 This modification increases the affinity of Polη for PCNA, with which it interacts via a single PCNA interacting peptide (PIP) box and a single ubiquitin-binding zinc finger (UBZ) domain.1,3In contrast to its yeast and mammalian counterparts, Polη from Arabidopsis and Oryza sativa (rice) has two PIP boxes and lacks a UBZ.6,9,10 Instead the two polymerases each possess two ubiquitin-binding motifs (UBMs) similar to those present in the Arabidopsis Rev1 protein and a vertebrate TLS polymerase, Pol., for which there is no homolog in Arabidopsis.6,13 Considerable differences in the sequences flanking the UBMs in Polη and Rev1 argue that Polη did not acquire its UBMs from Rev1, and so, although perhaps unique to plant Polη, their origin remains a mystery.The presence of PCNA- and ubiquitin-binding sequences in plant Polη hint that it may operate in TLS in a manner similar to that for Polη from yeast or mammalian cells. Indeed, three lines of evidence6 lead us to suggest that the Polη PIP boxes and UBMs likely function in binding ubiquitylated PCNA and this interaction is probably required for TLS past UV photoproducts by Arabidopsis Polη. First, Arabidopsis Polη interacts physically and in yeast two-hybrid assays with Arabidopsis PCNA1 and PCNA2. Second, expression in yeast of Arabidopsis cDNAs encoding Polη and PCNA2, but not PCNA1, fully complements the UV sensitivity conferred by elimination of yeast Polη. In vitro mutagenesis suggests the inability of Polη plus PCNA1 to restore UV resistance is due to a lysine at position 201 in PCNA1 but not PCNA2. In the three-dimensional structure of PCNA, amino acid 201 lies adjacent to lysine-164, the residue that is ubiquitylated in yeast and human PCNA. Thus, one possibility is that lysine-201 in PCNA1 prevents complementation of UV sensitivity by inhibiting ubiquitylation of lysine-164. Third, altering presumed critical residues in either of the two PIP boxes or UBM2 in Arabidopsis Polη also prevents restoration of UV resistance in Polη-deficient yeast cells.Several important parts of the puzzle remain to be solved. In particular, the ubiquitylation of plant PCNA has yet to be demonstrated, and the identity of the proteins that might monoubiquitylate plant PCNA is uncertain. Although Arabidopsis Rad6 homologs can ubiquitylate target proteins in vitro, there is no evidence that Arabidopsis PCNA1 or PCNA2 is a substrate, and Arabidopsis lacks a Rad18 homolog.14,15 Finally, if PCNA is ubiquitylated in planta, does this occur at lysine-164 in response to DNA damage or replication fork stalling, is the interaction of Polη with PCNA stimulated by this modification, and is an enhanced interaction mediated by the Polη UBMs?  相似文献   

16.
Plant defensins are small, highly stable, cysteine-rich peptides that constitute a part of the innate immune system primarily directed against fungal pathogens. Biological activities reported for plant defensins include antifungal activity, antibacterial activity, proteinase inhibitory activity and insect amylase inhibitory activity. Plant defensins have been shown to inhibit infectious diseases of humans and to induce apoptosis in a human pathogen. Transgenic plants overexpressing defensins are strongly resistant to fungal pathogens. Based on recent studies, some plant defensins are not merely toxic to microbes but also have roles in regulating plant growth and development.Key words: defensin, antifungal, antimicrobial peptide, development, innate immunityDefensins are diverse members of a large family of cationic host defence peptides (HDP), widely distributed throughout the plant and animal kingdoms.13 Defensins and defensin-like peptides are functionally diverse, disrupting microbial membranes and acting as ligands for cellular recognition and signaling.4 In the early 1990s, the first members of the family of plant defensins were isolated from wheat and barley grains.5,6 Those proteins were originally called γ-thionins because their size (∼5 kDa, 45 to 54 amino acids) and cysteine content (typically 4, 6 or 8 cysteine residues) were found to be similar to the thionins.7 Subsequent “γ-thionins” homologous proteins were indentified and cDNAs were cloned from various monocot or dicot seeds.8 Terras and his colleagues9 isolated two antifungal peptides, Rs-AFP1 and Rs-AFP2, noticed that the plant peptides'' structural and functional properties resemble those of insect and mammalian defensins, and therefore termed the family of peptides “plant defensins” in 1995. Sequences of more than 80 different plant defensin genes from different plant species were analyzed.10 A query of the UniProt database (www.uniprot.org/) currently reveals publications of 371 plant defensins available for review. The Arabidopsis genome alone contains more than 300 defensin-like (DEFL) peptides, 78% of which have a cysteine-stabilized α-helix β-sheet (CSαβ) motif common to plant and invertebrate defensins.11 In addition, over 1,000 DEFL genes have been identified from plant EST projects.12Unlike the insect and mammalian defensins, which are mainly active against bacteria,2,3,10,13 plant defensins, with a few exceptions, do not have antibacterial activity.14 Most plant defensins are involved in defense against a broad range of fungi.2,3,10,15 They are not only active against phytopathogenic fungi (such as Fusarium culmorum and Botrytis cinerea), but also against baker''s yeast and human pathogenic fungi (such as Candida albicans).2 Plant defensins have also been shown to inhibit the growth of roots and root hairs in Arabidopsis thaliana16 and alter growth of various tomato organs which can assume multiple functions related to defense and development.4  相似文献   

17.
Prion diseases are infectious conformational diseases. Despite the determination of many native prion protein (PrP) structures and in vitro production of infectious prions from recombinant PrP the structural background of PrP conversion remains the largest unsolved problem. The aggregated state of PrPSc makes it inaccessible to high resolution techniques, therefore indirect methods have to be used to investigate the conversion process. We engineered disulfide bridges into the structured domain of PrP in order to determine the secondary structure elements that remain conserved upon conversion. Rather surprisingly, introduction of disulfides into each or both of the subdomains B1-H1-B2 and H2-H3 of the C-terminal globular domain retained the robust ability to convert into fibrils with increased content of β-structure, indistinguishable from the wild-type PrP. On the other hand disulfide bridges tethering the two subdomains completely prevented conversion, while their reduction reversed their conversion ability. The same conversion propensity was replicated also in prion infected cell lines. Experiments with combinations of engineered cysteine residues further support that domain swapping, centered on the B2-H2 loop, previously associated to species barrier, leads to PrP swapped dimers as the building block of prion fibrils.Key words: PrP, prion protein, mPrP, murine prion protein, prion protein, structural conversion, disulfide crosslinks, secondary structure, domain swapping, rigid loop, dimerOur understanding of the molecular mechanisms of prion diseases recently significantly advanced with the invention of PMCA technique1 and the demonstration that the converted recombinant PrP can induce transmissible disease,25 which conclusively fulfills the Koch''s postulates of infectivity. Another important development, relevant to the structural background of conformational diseases was the demonstration of the ability of short peptides to form cross-β-structure in many diverse orientations forming the so called dry steric zippers, which might underlay the existence of different prion strains.6 However the main question on the biochemical and structural nature of PrP conversion process remained unanswered.Prion diseases are characterized by conversion of the native PrP into the form PrPSc, which forms amyloid aggregates that are resistant to proteolysis. Tertiary structure of the native form of PrP from more than 15 different species has been determined.7 Their fold is highly conserved, with an unstructured N-terminal half of the protein and a C-terminal structured domain consisting of three α-helices and two β-strands.8 The native form of PrP exhibits high content of α-helical structure, while the converted form is dominated by the β-type secondary structure. The secondary structure content of the PrPSc is somewhat controversial. Analyses of infrared or CD spectra suggest that the secondary structure of converted PrP contains between 17–30% of α-helix and 43–50% of β-structure.9,10 This is clearly different from the all-β structure. It is in fact compatible with the conservation of the large part of the secondary structure elements of the C-terminal globular domain and induced formation of the β-structure from the proximal N-terminal segment, disordered in the native state.The defined tertiary structure of proteins is determined by the multitude of cooperative interactions that provide the sufficient free energy gap between the native and nonnative conformations. The existence of alternative, significantly different global folds of any protein has not been demonstrated yet at the level of a defined tertiary structure. It would be in fact extremely difficult to stabilize the alternative stable fold, where many of the corporative interactions would have to be optimized simultaneously. This proposition is supported by the observation that most of the proteins involved in conformational diseases contain a segment with an intrinsically unfolded structure in the native state.11 Therefore it is much more likely for those unfolded segments to adopt an ordered conformation rather than to completely refold the native globular structure.Aggregation state of the PrPSc hinders determination of high resolution structure. We can however use different biochemical approaches to inquire about the nature of the conversion process and structure of the converted form. Methods, such as antibody mapping,12,13 hydrogen exchange,1416 binding of fluorescent ligands17 and many others have been used, revealing that both C-terminal and proximal N-terminal segment of the PrP become less accessible to the solvent upon conversion.In order to unravel the molecular mechanism of PrP conversion, we decided to investigate which of the secondary structure elements or their suprasecondary structure combinations are retained in the converted form. We introduced disulfide tethers into different positions within the globular C-terminal segment of mPrP, connecting different secondary structure elements.18 Several pairs of residues that adhere to the geometric requirements for a disulfide formation were selected.19 Covalent tethers impose a very strong structural constraint as the relative position of the tethered pair needs to remain the same in the converted structure. This approach therefore allowed us to probe the relative position of all secondary structure elements in the converted PrP. We successfully prepared seven disulfide-tethered variants of mPrP. The only variants that we could not prepare were those where the introduced cysteines were in the neighborhood of the existing disulfide and probably led to the heterogeneous disulfide shuffling yielding misfolded products. We demonstrated that in all variants additional disulfides are formed and the secondary structure of the native form of PrP variants is indistinguishable from the wild-type PrP.The key experiment was in vitro conversion of PrP disulfide mutants. Surprisingly, the majority of the disulfide-tethered variants was able to convert into PrP fibrils. Mutant fibrils had the same morphology, determined by AFM and TEM, pattern of antibody mapping and high content of β-structure as the fibrils prepared from wild-type PrP. The common structural property of the three variants that did not convert and retained the native, α-helical conformation, regardless of the conversion protocol, is that they all tethered the two subdomains B1-H1-B2 and H2-H3 to each other (Fig. 1). The proof that this is indeed an intrinsic structural property rather than a result of serendipitous point mutations is that both variants with single cysteine residues of the disulfide pair retained the ability to convert. Moreover reduction of disulfides rendered the originally non-converting disulfide variants convertible into the β-structured fibrils. Even simultaneous introduction of two new disulfides, one into each of the two subdomains, retained the ability of PrP variants to convert. Introduction of disulfides predominantly improved the stability of the protein, increasing the Tm by 3–12 degrees. However, the conversion ability had no correlation with the thermal stability of the protein as some of the most stable variants, containing two disulfides that increased Tm by more than 16 degrees, readily converted.Open in a separate windowFigure 1Mapping of PrP conversion by disulfide tethers. Disulfides engineered within the globular domain of PrP have different effects on its ability to convert into fibrils. Disulfide tethers are schematically represented as straight connectors on mouse PrP structure (1XYX).22 All disulfides (left top), which tether on one side subdomain B1-H1-B2 (gray) and on the other subdomain H2-H3 (black) prevent conversion, while PrP variants with single or even double disulfide tethers within each or both of the two subdomains retain the ability to convert into fibrils (left bottom). Results suggest that the secondary structure of each of the two subdomains is conserved during conversion, which can be accomplished by separation of subdomains (middle) followed by domain swapping. Domain swapped PrP dimer thus represents the building block of fibrils and a template for the annealing of the disordered N-terminal part into β-structure. Monomers within a swapped dimer are shown in gray and black (right).Those results provide an exceptionally strong set of constraints to characterize the conversion process and structure of the converted form. Our results are not compatible with most of the current structural models of PrP conversion, which suggest unfolding or significant rearrangements of secondary structure elements of the globular domain.14,16,20,21 Separation of subdomains of PrP implies that this process requires high activation energy or highly unfolding conditions. The loop linking the two subdomains connects B2 to H2 and has also been called “the rigid loop,” named by the increased ordering in the elk PrP in contrast to mouse or human PrP.22 This loop has been implicated in the species barrier23,24 and protective polymorphisms.25 Mice carrying mutations S170N N174T, where residues from mouse PrP are replaced with the corresponding residues from elk, develop spontaneous transmissible prion disease.26It might be in principle possible that disulfide variants convert off-pathway from the physiologically relevant PrPSc form. However we were able to demonstrate the same properties in cell cultures; only in vitro convertible PrP variant was able to replicate prions, while the unconvertible variant did not.The only structural transition that is compatible with our results is domain swapping of the C-terminal globular domain. Domain swapping represents the mechanism of oligomerization where the monomer and oligomer share the majority of the secondary structure elements. Most of the residues in the swapped-dimer oligomer are in exactly the same type of chemical environment as in the monomer with the exception of residues that represent the hinge of subdomain separation and connection between the monomeric units. Domain swapping requires high activation energy as the monomer has to unfold during conversion. The resulting oligomers are typically extremely stable and often a single protein can form different domain-swapped oligomers.27In order to confirm domain-swapped model of prion protein conversion we performed additional experiments where we analyzed the conversion products of a mixture of the two single cysteine mutants. Those single cysteine variants were designed in a way that if swapping of the sub-domains B1-H1-B2 and H2-H3 occurs during conversion, cysteines from different single cysteine variants come into contact and can form a disulfide bridge. Indeed proteinase K-resistant covalent dimers were only observed upon conversion of a mixture of both variants.In conclusion, we present the model of PrP conversion, where the conversion process requires unfolding of the core of the structured C-terminal domain of PrP with separation of the two subdomains, which recombine into a swapped dimer (Fig. 1). It has been demonstrated previously by several different approaches that PrP dimerization is important and a rate limiting step in conversion.2830 This swapped dimer represents the building block of fibrils and the template for structuring of the unfolded N-terminal segment, which can anneal to the dimer in the form of the β-strands, such as demonstrated in peptide dry steric zippers. We propose that the variability between different strains of prions may originate from differently annealed β-strands of the N-terminal segments and can additionally be affected by posttranslational modifications and the presence of additional molecules, such as nucleic acids or lipids.  相似文献   

18.
The 26S proteasome, a central enzyme for ubiquitin-dependent proteolysis, is a highly complex structure comprising 33 distinct subunits. Recent studies have revealed multiple dedicated chaperones involved in proteasome assembly both in yeast and in mammals. However, none of these chaperones is essential for yeast viability. PAC1 is a mammalian proteasome assembly chaperone that plays a role in the initial assembly of the 20S proteasome, the catalytic core of the 26S proteasome, but does not cause a complete loss of the 20S proteasome when knocked down. Thus, both chaperone-dependent and -independent assembly pathways exist in cells, but the contribution of the chaperone-dependent pathway remains unclear. To elucidate its biological significance in mammals, we generated PAC1 conditional knockout mice. PAC1-null mice exhibited early embryonic lethality, demonstrating that PAC1 is essential for mammalian development, especially for explosive cell proliferation. In quiescent adult hepatocytes, PAC1 is responsible for producing the majority of the 20S proteasome. PAC1-deficient hepatocytes contained normal amounts of the 26S proteasome, but they completely lost the free latent 20S proteasome. They also accumulated ubiquitinated proteins and exhibited premature senescence. Our results demonstrate the importance of the PAC1-dependent assembly pathway and of the latent 20S proteasomes for maintaining cellular integrity.The 26S proteasome is a eukaryotic ATP-dependent protease responsible for the degradation of proteins tagged with polyubiquitin chains (21). The ubiquitin-dependent proteolysis by the proteasome plays a pivotal role in various cellular processes by catalyzing the selective degradation of short-lived regulatory proteins as well as damaged proteins. Thus, the proteasome is essential for the viability of all eukaryotic cells.The 26S proteasome is a large protein complex consisting of two portions; one is the catalytic 20S proteasome of approximately 700 kDa (also called the 20S core particle), and the other is the 19S regulatory particle (RP; also called PA700) of approximately 900 kDa, both of which are composed of a set of multiple distinct subunits (70). The 20S proteasome is a cylindrically shaped stack of four heptameric rings, where the outer and inner rings each are composed of seven homologous α subunits (α1 to α7) and seven homologous β subunits (β1 to β7), respectively (5). The proteolytic active sites reside within the central chamber enclosed by the two inner β-rings, while a small channel formed by the outer α-ring, which is primarily closed, restricts the access of native proteins to the catalytic chamber. Thus, the 20S proteasome is a latent enzyme. Appending 19S RP, which consists of 19 different subunits, to the α-ring enables the 20S proteasome to degrade native proteins; 19S RP accepts ubiquitin chains of substrate proteins, removes ubiquitin chains while unfolding the substrates, and feeds the substrates into the interior proteolytic chamber of the 20S proteasome through the α-ring that is opened when the C-terminal tails of the ATPase subunits of 19S RP are inserted into the intersubunit spaces of the α-ring (24, 62, 74). However, it also has been reported that some denatured or unstructured proteins can be degraded directly by the 20S proteasome even in the absence of 19S RP and ubiquitination (37, 39).Much attention has been focused on how such a highly elaborate structure is achieved. Recent studies have identified various proteasome-dedicated chaperones that assist in the assembly of the proteasome in eukaryotic cells (23, 40, 56, 57, 65, 66). In yeast, while most of the proteasome subunits are essential for viability, the deletion of any of these chaperones does not cause lethality. In fact, many, if not all, of the deletions exhibit subtle phenotypes. In mammalian cells, although the knockdown of the assembly chaperones reduced proteasome assembly and thus proteasome activity, leading to slow cell growth, the degree of reduction was much lower than that which occurred following the knockdown of the proteasome subunit itself (33, 35, 40). These results indicate that the assembly chaperones play an auxiliary role in proteasome biogenesis.Proteasome assembly chaperone 1 (PAC1) is one of the assembly chaperones originally identified in mammalian cells (34). PAC1 plays a role in α-ring formation that occurs during the initial assembly of the 20S proteasome; it also prevents the aberrant dimerization of the α-ring. As is the case for most assembly chaperones, the knockdown of PAC1 in mammalian cells decreases proteasome activity but to a lesser extent than that in, for example, β2 knockdown (34, 35). Therefore, both PAC1-dependent and -independent assembly pathways exist in cells, but the importance of the PAC1-dependent pathway remains elusive. To further elucidate the biological significance of PAC1 and PAC1-dependent proteasome biogenesis, we generated conditional mouse mutants carrying an inactivating mutation in Psmg1, the gene coding for PAC1 protein, in the whole body, the nervous system, and in the liver. Our results demonstrate that PAC1 is essential for the development of a mouse, and that it plays important roles in maintaining cellular integrity in quiescent tissue. Our study revealed for the first time the importance of chaperone-mediated proteasome biogenesis in a whole-body mammalian system and may provide valuable knowledge in medical drug development targeting proteasomes.  相似文献   

19.
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   

20.
The discovery of new functions for platelets, particularly in inflammation and immunity, has expanded the role of these anucleate cell fragments beyond their primary hemostatic function. Here, four in-depth human platelet proteomic data sets were generated to explore potential new functions for platelets based on their protein content and this led to the identification of 2559 high confidence proteins. During a more detailed analysis, consistently high expression of the proteasome was discovered, and the composition and function of this complex, whose role in platelets has not been thoroughly investigated, was examined. Data set mining resulted in identification of nearly all members of the 26S proteasome in one or more data sets, except the β5 subunit. However, β5i, a component of the immunoproteasome, was identified. Biochemical analyses confirmed the presence of all catalytically active subunits of the standard 20S proteasome and immunoproteasome in human platelets, including β5, which was predominantly found in its precursor form. It was demonstrated that these components were assembled into the proteasome complex and that standard proteasome as well as immunoproteasome subunits were constitutively active in platelets. These findings suggest potential new roles for platelets in the immune system. For example, the immunoproteasome may be involved in major histocompatibility complex I (MHC I) peptide generation, as the MHC I machinery was also identified in our data sets.Although first described over a century ago, new roles and functions for platelets continue to emerge. Derived by budding from megakaryocytes and devoid of a nucleus, platelets were formerly not thought to produce proteins and their one role was to initiate and perform blood clotting. However, this view has changed in recent years; platelets have mRNA, microRNAs to regulate their mRNA, the machinery to synthesize proteins and they use it (1, 2). Furthermore, in addition to their function in hemostasis, it has been recognized that platelets play a role in inflammatory processes (3, 4). Through their interactions with the endothelium and other blood cells, platelets are believed to play a critical role in defense, wound repair, and more (5). Understanding of many of the new aspects of platelet function is still limited, but these recent advances raise the question of what other features are awaiting discovery that might be hidden in these small cell fragments.There are limited methods available with which to study platelets; DNA-based methods cannot be applied, and although mRNA is present in platelets, its low level only allows for restricted analysis. Mass spectrometry (MS)-based proteomics is particularly well set up to study platelets, and previous studies have analyzed the platelet proteome (611), various subproteomes (1216), and have shed light on aspects of platelet signaling and function (1721). In this study, proteomic analysis of human platelets was conducted, generating an inventory of platelet proteins, which was then explored by comparison to proteomic data sets of nucleated cells with the aim of identifying new biology-related functions. This approach revealed consistently high expression of the proteasome, the protein complex that is the main protein degradation machinery in cells (Fig. 1). The presence of the proteasome in platelets has been described earlier (22). It is known to be active and its activity increases in response to agonist stimulation (23); however, a detailed analysis of the many subunits of this multimeric complex has not been performed and its role in platelets, which produce less protein than nucleated cells, is not fully understood. The proteasome''s core complex, the 20S proteasome, is composed of 28 nonidentical subunits, arranged in four rings, two comprising of seven α subunits and two of seven β subunits. Three of the β subunits (β1, β2, and β5) are catalytically active. The 20S proteasome forms the 26S proteasome together with the 19S regulator, which contains ATPase subunits and is responsible for the ATP1 dependence of the 26S proteasome. The immunoproteasome, which is constitutively expressed in cells of the immune system or is synthesized following induction by interferon γ (IFNγ) in all other nucleated cells, is formed when the catalytically active β subunits are replaced by their immunoproteasome counterparts (β1i, β2i, and β5i). IFNγ also up-regulates the 11S regulator, which consists of PA28 α and β subunits, and both the immunoproteasome and the 11S proteasome are thought to be involved in improved peptide generation for major histocompatibility complex (MHC) I antigen presentation (24).Open in a separate windowFig. 1.Composition of the proteasome and immunoproteasome. The standard 20S core (middle) is composed of 28 nonidentical subunits that are arranged in four rings; two composed of seven α subunits and two composed of seven β subunits. Three of the β subunits (β1, β2, and β5) are catalytically active. The 19S regulator is composed of a base, containing six ATPase subunits and two non-ATPase subunits, and a lid, which contains up to ten non-ATPase subunits. The 20S proteasome and two 19S regulators form the 26S proteasome (left). The immunoproteasome, which is induced by IFNγ, contains three different catalytically active subunits (β1i, β2i, and β5i). The 11S regulator, which consists of heptameric complexes containing PA28α and β subunits, is also induced by IFNγ and can replace the 19S regulator (right).Here, discovery of the high expression of the proteasome in our platelet proteomic data set was followed up with traditional biochemical assays to explore in detail the composition of the proteasome in platelets. Not only were all components of the 26S proteasome detected in our global platelet data sets, but immunoproteasome subunits were also identified. We validated that all members of the 20S proteasome were present and assembled in human platelets. Furthermore, we show that the standard as well as the immunoproteasome catalytic subunits are active. The presence of not only active proteasome but active immunoproteasome subunits in platelets opens up the possibility of new roles for these anucleate players, and further illustrates the critical role proteomics plays in improving our understanding of platelet function.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

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