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.
Tens of putative interacting partners of the cellular prion protein (PrPC) have been identified, yet the physiologic role of PrPC remains unclear. For the first time, however, a recent paper has demonstrated that the absence of PrPC produces a lethal phenotype. Starting from this evidence, here we discuss the validity of past and more recent literature supporting that, as part of protein platforms at the cell surface, PrPC may bridge extracellular matrix molecules and/or membrane proteins to intracellular signaling pathways.Key words: prion protein, PrPC, extracellular matrix, cell adhesion molecules, neuritogenesis, p59fyn, Ca2+Initially, the discovery that the prion protein was the major, if not the unique, component of the prion agent causing transmissible spongiform encephalopathies (TSE)1 has placed the protein in an extremely unfavorable light. Thereafter, however, a wealth of evidence has supported the notion that the protein positively influences several aspects of the cell physiology, and that its duality—in harboring both lethal and beneficial potentials—could be rationalized in terms of a structural switch. Indeed, the protein exists in at least two conformational states: the cellular, α helix-rich isoform, PrPC, and the prion-associated β sheet-rich isoform, PrPSc.2 If it is now unquestionable that the presence of PrPC in the cell is mandatory for prion replication and neurotoxicity to occur,3,4 nonetheless its physiologic function is still debatable, despite the long lasting effort, and the numerous, frequently genetically advanced, animal and cell model systems dedicated to the issue. From these studies the picture of an extremely versatile protein has emerged, whereby PrPC acts in the cell defense against oxidative and apoptotic challenges, but also in cell adhesion, proliferation and differentiation, and in synaptic plasticity.5,6 In an effort to converge these multiple propositions in an unifying functional model, different murine lines devoid of PrPC have been studied. These animals, however, displayed no obvious phenotype,79 suggesting that either PrPC is dispensable during development and adult life or that compensative mechanisms mask the loss of PrPC function in these paradigms. Thus, identifying the exact role of PrPC in the cell would not only resolve an important biological question, but would also help elucidate the cellular steps of prion pathogenesis necessary for designing early diagnostic tools and therapeutic strategies for TSE.As is often the case, the employment of a model system unprecedented in prion research has recently disclosed a most interesting scenario with regards to PrPC physiology, having unravelled, for the first time, a lethal phenotype linked to the absence of the protein.10 The paradigm is the zebrafish, which expresses two PrPC isoforms (PrP1 and PrP2). Similarly to mammalian PrPC, they are glycosylated and attached to the external side of the plasma membrane through a glycolipid anchor. PrP1 and PrP2 are, however, expressed in distinct time frames of the zebrafish embryogenesis. Accordingly, the knockdown of the PrP1, or PrP2, gene very early in embryogenesis impaired development at different stages, bypassing putative compensatory mechanisms. By focusing on PrP1, Malaga-Trillo et al. showed that the protein was essential for cell adhesion, and that this event occurred through PrP1 homophilic trans-interactions and signaling. This comprised activation of the Src-related tyrosine (Tyr) kinase p59fyn, and, possibly, Ca2+ metabolism, leading to the regulation of the trafficking of E-cadherin, a member of surface-expressed cell adhesion molecules (CAMs) responsible for cell growth and differentiation.11 It was also reported that overlapping PrP1 functions were performed by PrPCs from other species, while the murine PrPC was capable to replace PrP1 in rescuing, at least in part, the knockdown developmental phenotype. Apart from providing the long-sought proof for a vital role of PrPC, the demonstration that a mammalian isoform corrected the lethal zebrafish phenotype strongly reinforces previous results—mainly obtained in a variety of mammalian primary neurons and cell lines—pointing to a functional interplay of PrPC with CAMs, or extra cellular matrix (ECM) proteins, and cell signaling, to promote neuritogenesis and neuronal survival. A revisit of these data is the main topic of the present minireview.As mentioned, the capacity of PrPC to act as a cell adhesion, or recognition, molecule, and to entertain interactions with proteins implicated in growth and survival, has already been reported for the mammalian PrPC. A case in point is the interaction, both in cis- and trans-configurations, with the neuronal adhesion protein N-CAM12 that led to neurite outgrowth.13 Like cadherins, N-CAM belongs to the CAM superfamily. Following homo- or heterophylic interactions, it can not only mediate adhesion of cells, or link ECM proteins to the cytoskeleton, but also act as a receptor to transduce signals ultimately resulting in modulating neurite outgrowth, neuronal survival and synaptic plasticity.11 Another example is the binding of PrPC to laminin, an ECM heterotrimeric glycoprotein, which induced neuritogenesis together with neurite adhesion and maintenance,14,15 but also learning and memory consolidation.16 Further, it has been described that PrPC interacted with the mature 67 kDa-receptor (67LR) (and its 37 kDa-precursor) for laminin, and with glycosamminoglycans (GAGs), each of which is involved in neuronal differentiation and axon growth.1721 More recently, Hajj et al.22 have reported that the direct interaction of PrPC with another ECM protein, vitronectin, could accomplish the same process, and that the absence of PrPC could be functionally compensated by the overexpression of integrin, another laminin receptor.23 Incidentally, the latter finding may provide a plausible explanation for the absence of clear phenotypes in mammalian PrP-null paradigms. By exposing primary cultured neurons to recombinant PrPs, others have shown that trans-interactions of PrPC are equally important for neuronal outgrowth,24,25 including the formation of synaptic contacts.25 Finally, it has been demonstrated that the binding of PrPC with the secreted co-chaperone stress-inducible protein 1 (STI1) stimulated neuritogenesis.26 This same interaction had also a pro-survival effect, as did the interaction of PrPC with its recombinant form.24 Notably, the involvement of PrPC in cell protection has been heightened by experiments with whole animals. By applying transient or permanent focal cerebral ischemia to the animals, it was found that their reduced brain damage correlated with spontaneous or adenoviral-mediated, upregulation of PrPC,2729 (reviewed in ref. 30), and that PrPC deficiency aggravated their ischemic brain injury.30,31 Thus, now that data are available from phylogenetically distant paradigms (zebrafish and mammalian model systems), it acquires more solid grounds the advocated engagement of PrPC in homo/heterophilic cis/trans interactions to trigger signaling events aiming at neuronal—or, in more general terms, cell—survival and neuritogenesis. The latter notion is consistent with the delayed maturation of different types of PrPC-less neurons, observed both in vitro and in vivo.32,33If one assumes that the interaction of PrPC with multiple partners (45 for PrPC and PrPSc, as reviewed in Aguzzi et al.,5 or 46 considering the homophylic interaction) are all functionally significant, the most immediate interpretation of this “sticky” behavior entails that PrPC acts as a scaffolding protein in different membrane protein complexes.5,6 Each complex could then activate a specific signaling pathway depending on the type and maturation of cells, the expression and glycosylation of PrPC, and availability of extra- and intra-cellular signaling partners. At large, all these signals have been shown to be advantageous to the cell. However, because in a cell only a subtle line divides the “good” from the “bad,” instances can be envisioned in which a pro-life signal turns into a pro-death signal. A typical example of this possibility is glutamate excitotoxicity resulting in dangerous, glutamate receptor-linked, Ca2+ overload. Likewise, an excessive or over-stimulated signal elicited by PrPC, or by the putative complex housing the protein could become noxious to the cell. This possibility may explain why the massive expression of PrPC caused degeneration of the nervous system,34 and of skeletal muscles,34,35 in transgenic animals. More intriguing is the finding that—in a mouse line expressing anchorless PrPC—PrPSc was capable to replicate without threatening the integrity of neurons.36 This may suggest that native membrane-bound PrPC acts as, or takes part into, a “receptor for PrPSc”, and that lasting PrPSc-PrPC interactions distort the otherwise beneficial signal of the protein/complex and cause neurodegeneration.37 Consistent with this hypothesis is the finding that the in vivo antibody-mediated ligation of PrPC provoked apoptosis of the antibody-injected brain area.38 Speculatively, the action of N-terminally, or N-proximally truncated PrPs whose expression in PrP-less transgenic mice induced extensive neurodegeneration,3941 may be traced back to the same hyper-activation of PrPC signaling. Possibly, this may hold true also for the synaptic impairment that, recorded only in PrPC-expressing neurons, was attributed to the binding of amyloid beta (Aβ) peptide oligomers implicated in Alzheimer disease, to PrPC.42,43But which is (are) the cellular signaling pathway(s) conveyed by the engagement of PrPC in different signaling complexes? In line with its multifaceted behavior, several intracellular effectors have been proposed, including p59fyn, mitogen-activated kinases (MAPK) Erk1/2, PI3K/Akt and cAMP-PKA. p59fyn is the most reported downstream effector, suggesting that, in accordance with its behavior, p59fyn could serve as the sorting point for multiple incoming and outgoing signals also in the case of PrPC. The initial evidence of the PrPC-p59fyn connection came from cells subjected to antibody-mediated cross-linking of PrPC.44 Later, it was shown that the PrPC-p59fyn signal converged to Erk1/2 through a pathway dependent on (but also independent of) reactive oxygen species generated by NADPH oxidase.45 A PrPC-dependent activation of p59fyn13,25 and Erk1/2 (but also of PI3K and cAMP-PKA)24 was evident in other neuronal cell paradigms and consistent with the almost ubiquitous expression of PrPC, in non-neuronal cells such as Jurkat and T cells.46 Not to forget that in zebrafish embryonic cells activated p59fyn was found in the same focal adhesion sites harboring PrP1.10 Regarding the activation of the ERK1/2 pathway promoted by the PrPC-STI1 complex, and leading to neuritogenesis, the role of p59fyn was not investigated.26 The same holds true for the transduction of a neuroprotective signal by the PrPC-STI1 complex involving the cAMP-PKA pathway.26 Interestingly, this is not the only example reporting that engagement of PrPC activates simultaneously two independent pathways. In fact, possibly after transactivating the receptor for the epidermal growth factor, the antibody-mediated clustering of PrPC was shown to impinge on both the Erk1/2 pathway, and on a protein (stathmin) involved in controlling microtubule dynamics.47Yet, if p59fyn is implicated in mammalian PrPC-activated signaling cascade, a protein linking extracellular PrPC to p59fyn is needed, given the attachment of the enzyme to the inner leaflet of the plasma membrane through palmitoylated/myristoylated anchors. In this, the PrPC partner N-CAM (isoform 140) seems ideal to fulfill the task, given that p59fyn is part of N-CAM-mediated signaling. Indeed, after recruitment of N-CAM to lipid rafts—which may also depend on PrPC,13—together with the receptor protein Tyr phosphatase α (RPTPα), the Tyr-phosphate removing activity of RPTPα allows the subsequent activation of p59fyn through an autophosphorylation step.48 This event recruits and activates the focal adhesion kinase (FAK),11 another non-receptor Tyr kinase. Finally, formation of the FAK-p59fyn complex triggers neuritogenesis through both Erk1/2 and PI3K/Akt pathways.49,50 Parenthetically, the FAK-p59fyn and PI3K/Akt connection would be suitable to explain why aggravation of ischemic brain injury in PrP-deficient brains was linked to a depressed Akt activation.31 FAK-p59fyn complex, however, may be also involved in the signal triggered by the still mysterious PrPC partner, 67LR. This protein was reported not only to act as a laminin receptor but also to facilitate the interaction of laminin with integrins,51 thereby possibly activating (through integrins) FAK-p59fyn-regulated pathways.49 Conversely, other data have supported the candidature of caveolin-1 for coordinating the signal that from PrPC reaches Erk1/2 through p59fyn.44,45,52 Further scrutiny of this route has shown that it comprised players such as laminin and integrins (upstream), FAK-p59fyn, paxillin and the Src-homology-2 domain containing adaptor protein (downstream), and that caveolin-1, a substrate of the FAK-p59fyn complex, facilitated the interaction of these signaling partners by recruiting them in caveolae-like membrane domains.53For the relevance they bear, we need to acknowledge recent propositions supporting the commitment of PrPC with proteins whose function is unrelated from the above-mentioned cell adhesion or ECM molecules; namely, the β-site amyloid precursor protein (APP) cleaving enzime (BACE1) and the N-methyl-D-aspartate (NMDA)-receptor. BACE1 is a proteolytic enzyme involved in Aβ production. It has been shown that overexpressed PrPC restricted, while depletion of PrPC increased the access of BACE1 to APP, possibly because PrPC interacts with BACE1 via GAGs.54 Thus, native PrPC reduces the production of Aβ peptides. A beneficial effect of PrPC was also highlighted by Khosravani et al.55 showing that, by physically associating with the subunit 2D of the NMDA-receptor, PrPC attenuated neuronal Ca2+ entry and its possible excitotoxic effect. This clear example for the control of PrPC on Ca2+ metabolism is particularly intriguing in light of previous reports linking Ca2+ homeostasis to PrPC pathophysiology (reviewed in ref. 56). Also, it is important to mention that a few partners of PrPC or downstream effectors may initiate signals that increase intracellular Ca2+, and that, in turn, local Ca2+ fluctuations regulate some of the afore-mentioned pathways.11,49,57,58In conclusion, although still somehow speculative, the implication of Ca2+ in PrPC-dependent pathways raises the possibility that the different input signals originating from the interaction of PrPC with diverse partners may all converge to the universal, highly versatile Ca2+ signaling. Were indeed this the case, then clearly the acting of PrPC as Harlequin, the famous character of the 18th century Venetian playwright Carlo Goldoni, who struggles to fill the orders of two masters, would be merely circumstantial.  相似文献   

2.
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   

3.
4.
The “protein only” hypothesis states that the key phenomenon in prion pathogenesis is the conversion of the host protein (PrPC) into a β-sheet enriched polymeric and pathogenic conformer (PrPSc). However the region of PrP bearing the information for structural transfer is still controversial. In a recent report, we highlighted the role of the C terminal part i.e., the helixes H2 and H3, using mutation approaches on recombinant PrP. The H2H3 was shown to be the minimal region necessary to reproduce the oligomerization pattern of the full-length protein. The oligomers produced from isolated H2H3 domain presented the same structural characteristics as the oligomers formed from the full-length PrP. Combining other groups'' results, this paper further discusses the relative, direct or indirect role of different PrP regions in assembly. The H2H3 region represents the core of PrP oligomers and fibrils, whereas the N terminus could explain divergences among different aggregates. Finally this review evocates the possibility to separate the domain involved in prion information transference (i.e., prion replication) from the domain bearing the cytotoxicity properties.Key words: prion, H2H3, amyloid, domain of replication, unfolding, strain, polymer, fibersTransmissible spongiform encephalopathies (TSE), fatal neurodegenerative diseases affecting humans and other mammalians, induce in most cases loss of motor control and dementia. PrP is a protein physiologically present in parts of the animal kingdom (in mammals, birds, reptiles and fishes). According to the “protein-only” hypothesis,1,2 the key phenomenon in the pathogenesis is the conversion of the α-helix rich host-encoded PrP form (PrPC) into a pathogenic conformer (PrPSc) characterized by a higher content in β-sheet and a polymeric state. The conversion to an enriched β-sheet structure is supposed to be due to the modification—induced only by a PrPSc-like state acting as a template- of PrPC into the PrPSc conformer. This hypothesis was first proposed by Griffith in 19673 and revisited by Lansbury et al. in 1993.4 The prion hypothesis has now found increasing support from experimental evidence based on the synthetic production of β-sheeted recombinant PrP which shows pathogenic properties in a wide variety of physico-chemical conditions.57 However, the molecular basis of prion conversion remains unclear, especially the various structural landscape of the PrPSc, which is the basis of the strain phenomenon.8To understand the mechanisms of transfer of the structural information, two mains issues have to be addressed: (1) we need to understand which region(s) of the protein act as template for conversion and (2) what is the “pathogenic” state of this domain. In this review, we shall assume that the region bearing the infectious information for replication and the region responsible for polymerization are identical. However, the link between the propensity of a domain to form aggregates and the ability to contain the necessary information for prion replication is far from being trivial. Generally the formation of amyloid assemblies results from the aggregation of disordered peptides or in some cases from disordered regions of a folded protein.8 If we consider that prion replication is only supported by the globular part of PrP9 the currently available model involves the folded domain. Since all structural transitions need at least a partial unfolding and refolding process, pre-required structural events should be considered prior to the conversion process.  相似文献   

5.
There is increasing evidence that cellular prion protein plays important roles in neurodegeneration and neuroprotection. One of the possible mechanism by which this may occur is a functional inhibition of ionotropic glutamate receptors, including N-Methyl-D-Aspartate (NMDA) receptors. Here we review recent evidence implicating a possible interplay between NMDA receptors and prions in the context of neurodegenerative disorders. Such is a functional link between NMDA receptors and normal prion protein, and therefore possibly between these receptors and pathological prion isoforms, raises interesting therapeutic possibilities for prion diseases.Key words: NMDA, NR2D, glutamate, neuroprotection, calciumPrions are most often discussed in the context of transmissible spongiform encephalopathies (TSEs) which encompass a range of neurological disorders that include human Creutzfeldt-Jakob disease (among others), sheep scrapie and bovine spongiform encephalopathy.1,2 It is well established that these disorders arise from a progressive conversion of the normal, mainly helical form of cellular prion protein (PrPC) into a different PrPSc protein conformation with a high beta sheet content.3 In their PrPSc form, prions act as templates that catalyze misfolding of PrPC to produce increasing levels of PrPSc, which likely represents several or even many different conformational states of the same source protein, resulting in diverse clinical phenotypes. This in turn leads to accumulation of PrPSc deposits in the brain that can appear as aggregates and amyloid-like plaques4 and which disrupt normal neurophysiology.5 While the neuropathology of TSE''s has been explored in great detail dating back to the 1920s,6 less effort has perhaps been expended on understanding the cellular and physiological function of PrPC which is ubiquitously expressed, and found even in simple organisms.5,7,8 A number of mouse lines either lacking PrPC or overexpressing PrPC have been created, including the widely used Zurich I PrPC knockout strain.9,10 Despite the wide distribution of PrPC in the mammalian CNS, it perhaps surprisingly has only a relatively mild behavioral phenotype that appears to include some deficits in spatial learning at the behavioral level11,12 as well as alterations in long term potentiation at the cellular level.1317 In addition, it has been shown that these mice show an increased excitability of hippocampal neurons.13,1820 In contrast, deletion of certain parts of the PrPC protein in vivo can have serious physiological consequences. For example, deletion of a stretch of amino acids between just upstream of the octarepeat copper binding motifs produces a lethal phenotype, that can be rescued by overexpression of increasing levels of normal PrPC.21,22 Of particular note, these deletion mutants show degeneration of axons and myelin, both in the CNS and in peripheral nerves; indeed some mutants show a predilection for axomyelinic degeneration with little neuronal pathology,21 suggesting that certain mutated forms of PrP have a direct toxic effect on oligodendrocytes and/or myelin.23 Moreover, activation of the Dpl1 gene in mice lacking PrPC leads to an ataxic phenotype, that is not observed in the presence of PrPC.24 Collectively, this indicates that PrPC may act in a protective capacity and in contrast, certain abnormal forms of PrP are “toxic”, promoting much more injury to various elements of the CNS and PNS than outright absence of wild-type PrPC.This notion is further corroborated by a number of studies in PrPC knockout mice, both in vivo and in cell culture models. Cultured hippocampal neurons from PrPC null mice display greater apoptosis during oxidative stress.25 Moreover, overexpression of PrPC in rats protects them from neuronal damage during ischemic stroke, whereas PrPC null mice show greater damage.2729 When PrPC null mice are subjected to different types of seizure paradigms, they showed increased mortality and increased numbers of seizures.30 This increased neuronal damage can be diminished by the NMDA receptor blocker MK-801,31 potentially implicating glutamate receptors in this process. Finally, it was recently shown that the absence of PrPC protein protects neurons from the deleterious effects of beta amyloid, a protein involved in Alzheimer disease.32 It is important to note that NMDA receptors have been implicated in seizure disorders and in cell death during ischemic stroke.3335 Indeed, our own work has shown that NMDA receptors expressed endogenously in myelin contribute to myelin damage and may be one of the first steps leading to demyelination.36 Furthermore, the NMDA receptor blocker memantine is used to treat Alzheimer disease, implicating NMDA receptors. The observations above suggest that there may be an interplay between NMDA receptor activity and the physiological function of PrPC. In support of this hypothesis, our recent work has directly identified a common functional and molecular link between NMDA receptors and PrPC.37 Brain slices obtained from Zurich I PrPC null mice showed an increased excitability of hippocampal slices, which could be ablated by blocking NMDA receptor activity with amino-5-phosphonovaleric acid. Removal of extracellular magnesium ions to enhance NMDA receptor activity resulted in stronger pro-excitatory effects in slices and cultured neurons from PrPC null mice compared with those from normal animals. Synaptic recordings indicate that the amplitude and duration of NMDA mediated miniature synaptic currents is increased in PrPC null mouse neurons, and evoked NMDA receptor currents show a dramatic slowing of deactivation kinetics in PrPC null mouse neurons. The NMDA current kinetics observed in these neurons were qualitatively consistent with NMDA receptors containing the NR2D subunit.38 Consistent with a possible involvement of NR2D containing receptors, siRNA knockdown of NR2D normalized current kinetics in PrP-null mouse neurons. Furthermore, a selective co-immunoprecipitation between PrPC and the NR2D, but not NR2B subunits, was observed. This then may suggest the possibility that under normal circumstances, PrPC serves to suppress NR2D function, but when PrPC is absent, NR2D containing receptors become active, and because of their slow kinetics, may contribute to calcium overload under circumstances where excessive (or even normal) levels of glutamate are present. This would include conditions such as epileptic seizures, ischemia and Alzheimer disease, thus providing a possible molecular explanation for the link between PrPC and neuroprotection under pathophysiological conditions. Indeed, NMDA promoted greater toxicity in PrPC null mouse neurons, and upon injection into brains of PrPC null mice. It is interesting to note that one of the major NMDA receptor subtypes expressed in myelin is NR2D, thus bridging the observations of Micu et al.36 of NMDA receptor mediated cell death in ischemic white matter, and those of Baumann and colleagues21 showing that PrPC deletion mutants can cause damage to myelin.How might PrPC deletion mutants affect neuronal survival? One possibility may be that these deletion mutants compete with normal PrPC for NMDA receptors, but are unable to functionally inhibit them. Alternatively, it is possible that the PrPC deletion mutants, by virtue of binding to the receptors, may in fact increase receptor activity, thus causing increased cell death. In both cases, increasing the expression of normal PrPC would be expected to outcompete the deletion variants, thus reestablishing the protective function. A similar mechanism could perhaps apply to TSEs. It is possible that the PrPSc form, perhaps in a manner reminiscent of the PrPC deletion mutants, may be unable to inhibit NMDAR function, or perhaps would even enhance it. Any excess glutamate that may be released as a result of cell damage due to PrPSc aggregates, or even normally released amounts glutamate during the course of physiological neuronal signaling, could be sufficient to cause NMDAR mediated cell death and neuronal degeneration. In this context, it is interesting to note that chronic administration of the weakly NR2D selective inhibitor memantine delays death as a consequence of scrapie infection in mice.39 In the context of Alzheimer disease, binding of PrPC to beta amyloid may prevent the inhibitory action of PrPC on NMDA receptor function, thus increasing NMDA receptor activity and promoting cell death. This then may perhaps explain the beneficial effects of memantine in the treatment of Alzheimer disease.In summary, despite the fact that PrPC is one of the most abundantly expressed proteins in the mammalian CNS, its physiological role is uncertain. Recent observations from our labs have established an unequivocal functional link between normal prion protein and the ubiquitous excitatory NMDA receptor. Thus, one of the key physiological roles of PrPC may be regulation of NMDA receptor activity. The presence of abnormal species of prion protein, whether acquired via “infection”, spontaneous conformational conversion or genetically inherited, may in turn alter normal function and regulation of NMDA receptors, leading to chronic “cytodegeneration” of elements in both gray and white matter regions of the CNS. This key functional link between PrP and glutamate receptors may provide our first opportunity for rational therapeutic design against the devastating spongiform encephalopathies and potentially other neurodegenerative disorders not traditionally considered as TSE''s.  相似文献   

6.
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   

7.
8.
9.
Diabetes mellitus type 2 (DM2) results from the combination of insulin unresponsiveness in target tissues and the failure of pancreatic β cells to secrete enough insulin.1 It is a highly prevalent chronic disease that is aggravated with time, leading to major complications, such as cardiovascular disease and peripheral and ocular neuropathies.2 Interestingly, therapies to improve glucose homeostasis in diabetic patients usually involve the use of glibenclamide, an oral hypoglycemic drug that blocks ATP-sensitive K+ channels (KATP),3,4 forcing β cells to release more insulin to overcome peripheral insulin resistance. However, sulfonylureas are ineffective for long-term treatments and ultimately result in the administration of insulin to control glucose levels.5 The mechanisms underlying β-cell failure to respond effectively with glibenclamide after long-term treatments still needs clarification. A recent study demonstrating that this drug activates TRPA1,6 a member of the Transient Receptor Potential (TRP) family of ion channels and a functional protein in insulin secreting cells,7,8 has highlighted a possible role for TRPA1 as a potential mediator of sulfonylurea-induced toxicity.  相似文献   

10.
11.
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   

12.
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   

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.
Prion protein (PrP)-like molecule, doppel (Dpl), is neurotoxic in mice, causing Purkinje cell degeneration. In contrast, PrP antagonizes Dpl in trans, rescuing mice from Purkinje cell death. We have previously shown that PrP with deletion of the N-terminal residues 23–88 failed to neutralize Dpl in mice, indicating that the N-terminal region, particularly that including residues 23–88, may have trans-protective activity against Dpl. Interestingly, PrP with deletion elongated to residues 121 or 134 in the N-terminal region was shown to be similarly neurotoxic to Dpl, indicating that the PrP C-terminal region may have toxicity which is normally prevented by the N-terminal domain in cis. We recently investigated further roles for the N-terminal region of PrP in antagonistic interactions with Dpl by producing three different types of transgenic mice. These mice expressed PrP with deletion of residues 25–50 or 51–90, or a fusion protein of the N-terminal region of PrP with Dpl. Here, we discuss a possible model for the antagonistic interaction between PrP and Dpl.Key words: prion protein, doppel, neurotoxic signal, neurodegeneration, neuroprotection, prion diseaseThe normal prion protein, termed PrPC, is a membrane glycoprotein tethered to the outer cell surface via a glycosylphosphatidylinositol (GPI) anchor moiety.1,2 It is ubiquitously expressed in neuronal and non-neuronal tissues, with highest expression in the central nervous system, particularly in neurons.3 The physiological function of PrPC remains elusive. We and others have shown that PrPC functionally antagonizes doppel (Dpl), a PrP-like GPI-anchored protein with ∼23% identity in amino acid composition to PrP, protecting Dpl-induced neurotoxicity in mice.47 Dpl is encoded on Prnd located downstream of the PrP gene (Prnp) and expressed in the testis, heart, kidney and spleen of wild-type mice but not in the brain where PrPC is actively expressed.4,5,8 However, when ectopically expressed in brains, particularly in cerebellar Purkinje cells, Dpl exerts a neurotoxic activity, causing ataxia and Purkinje cell degeneration in Ngsk, Rcm0 and Zrch II lines of mice devoid of PrPC (Prnp0/0).4,9,10 In these mice, Dpl was abnormally controlled by the upstream Prnp promoter.4,5 This is due to targeted deletion of part of Prnp including a splicing acceptor of exon 3.11 Pre-mRNA starting from the residual exon1/2 of Prnp was abnormally elongated until the end of Prnd and then intergenically spliced between the residual Prnp exons 1/2 and the Prnd coding exons.4,5 As a result, Dpl was ectopically expressed under the control of the Prnp promoter in the brain, particularly in neurons including Purkinje cells.4,5 In contrast, in other Prnp0/0 lines, such as Zrch I and Npu, the splicing acceptor was intact, resulting in normal Purkinje cells without ectopic expression of Dpl in the brain.4The molecular mechanism of the antagonistic interaction between PrPC and Dpl remains unknown. We recently showed that the N-terminal half of PrPC includes elements that might mediate cis or trans protection against Dpl in mice, ameliorating Purkinje cell degeneration.12 We also showed that the octapeptide repeat (OR) region in the N-terminal domain is dispensable for PrPC to neutralize Dpl neurotoxicity in mice.12 Here, possible molecular mechanisms for the antagonism between PrPC and Dpl will be discussed.  相似文献   

16.
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   

17.
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   

18.
19.
20.
The serpins are the largest superfamily of protease inhibitors. They are found in almost all branches of life including viruses, prokaryotes and eukaryotes. They inhibit their target protease by a unique mechanism that involves a large conformational transition and the translocation of the enzyme from the upper to the lower pole of the protein. This complex mechanism, and the involvement of serpins in important biological regulatory processes, makes them prone to mutation-related diseases. For example the polymerization of mutant α1-antitrypsin leads to the accumulation of ordered polymers within the endoplasmic reticulum of hepatocytes in association with cirrhosis. An identical process in the neuron specific serpin, neuroserpin, results in the accumulation of polymers in neurons and the dementia FENIB. In both cases there is a clear correlation between the molecular instability, the rate of polymer formation and the severity of disease. A similar process underlies the hepatic retention and plasma deficiency of antithrombin, C1 inhibitor, α1-antichymotrypsin and heparin co-factor II. The common mechanism of polymerization has allowed us to group these conditions together as a novel class of disease, the serpinopathies.Key Words: serpins, α1-antitrypsin, neuroserpin, polymerization, dementia, conformational disease, serpinopathiesSerpins (or serine protease inhibitors) are the largest family of protease inhibitors. They have been found in all major branches of life including viruses, prokaryotes and eukaryotes.13 Despite their name there is increasing evidence that serpins can also inhibit other classes of proteases as demonstrated by the viral serpin CrmA and recently by a plant serpin, serpin1.4,5 They can even play a non-inhibitory role in events as diverse as blood pressure regulation (angiotensinogen), chromatin condensation (MENT), tumor progression (maspin), protein folding (hsp47) and hormone transport (cortisol and thyroxine binding globulin).6One of the most important roles of serpins is the regulation of enzymes involved in proteolytic cascades. Among these serpins are α1-antitrypsin, α1-antichymotrypsin, C1 inhibitor, antithrombin and plasminogen activator inhibitor-1, which play an important role in the control of proteases involved in the inflammatory, complement, coagulation and fibrinolytic pathways, respectively.1,3 The serpin superfamily is characterised by more than 30% homology with the archetypal serpin α1-antitrypsin and conservation of tertiary structure.7,8 Serpins adopt a metastable conformation composed in most cases of 9 α-helices, three β-sheet (A to C) and an exposed mobile reactive centre loop (RCL). This flexible RCL typically contains 20 residues that act as a pseudo substrate for the target protease (Fig. 1A).915 After formation of a Michaelis complex16,17 the enzyme cleaves the P1-P1′ bond of the serpin, releasing the P1'' residue and forming an ester bond between the protease and the serpin.18,19 This is then followed by a dramatic conformational transition from a stressed to relaxed conformation with the enzyme being pulled from the upper to the lower pole of the serpin and the insertion of the reactive loop as an extra strand in β-sheet A.2025 As a consequence of this conformational change the thermal stability of the serpin is greatly enhanced. Whereas a typical serpin in its native state exhibits a midpoint of thermal denaturation of around 50–60°C, a cleaved serpin with its RCL fully incorporated into β-sheet A denatures at temperatures >120°C.9,26,27 Another consequence is the inactivation of the enzyme, stabilised at the acyl-intermediate and unable to proceed further to deacylation of the complex.24,28 This serpin-protease complex then binds to members of the lipoprotein receptor family and is cleared from the circulation.2931Open in a separate windowFigure 1Inhibition of neutrophil elastase by α1-antitrypsin and the structural basis of polymerization. (A) After docking (left) the neutrophil elastase (grey) is inactivated by movement from the upper to the lower pole of the protein (right). This is associated with the insertion of the RCL (red) as an extra strand into β-sheet A (green). (B) The structure of α1-antitrypsin is centred on β-sheet A (green) and the mobile reactive centre loop (red). Polymer formation results from the Z variant of α1-antitrypsin (Glu342Lys at P17; indicated by arrow) or mutations in the shutter domain (blue circle) that open β-sheet A to favour partial loop insertion and the formation of an unstable intermediate (M*). The patent β-sheet A then accepts the loop of another molecule to form a dimer (D), which then extends into polymers (P). The individual molecules of α1-antitrypsin within the polymer, although identical, are coloured red, yellow and blue for clarity. Figure reproduced with permission from Lomas et al.97Despite the evolutionary advantage conferred upon serpins by the remarkable mobility of the native state, their complexity is also their weak point.19,32 Mutations affecting the serpins can lead to a variety of diseases, resulting from either a gain or loss of function.6,19 For example mutations can cause aberrant conformational transitions that result in the retention of the serpin within the cell of synthesis. This will lead to either protein overload and death of the cell in which the serpin is synthesised, or disease as a consequence of the resulting plasma deficiency. Such a mechanism underlies diseases as diverse as cirrhosis, thrombosis, angio-oedema, emphysema and dementia. We review here the common mechanism underlying these diseases that we have grouped together as the serpinopathies.3335 The aggregation and accumulation of conformationally destabilized proteins is an important feature of many neurodegenerative diseases, including Alzheimer''s and Parkinson''s disease and the spongiform encephalopathies. Indeed we have used the serpinopathies as a paradigm for these other ‘conformational diseases’.36  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