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 62 毫秒
1.
Understanding the mechanism by which prion infectivity is encoded by the misfolded protein PrPSc remains a high priority within the prion field. Work from several groups has indicated cellular cofactors may be necessary to form infectious prions in vitro. The identity of endogenous prion conversion cofactors is currently unknown, but may include polyanions and/or lipid molecules. In a recent study, we manufactured infectious hamster prions containing purified PrPSc, co-purified lipid and a synthetic photocleavable polyanion. The polyanion was incorporated into infectious PrPSc complexes and then specifically degraded by exposure to ultraviolet light. Light-induced in situ degradation of the incorporated polyanion had no effect on the specific infectivity of the samples as determined by end-point dilution sPMCA and scrapie incubation time assays. Furthermore, prion strain properties were not changed by polyanion degradation, suggesting that intact polyanions are not required to maintain the infectious properties of hamster prions. Here, we review these results and discuss the potential roles cofactors might play in encoding prion infectivity and/or strain properties.Key words: prion, polyanion, photodegradation, incorporation, PrPThe prion diseases are infectious diseases that are believed to be caused by the conformational change of a host-encoded protein, PrPC, to a pathogenic conformer PrPSc. The controversial “protein-only” hypothesis posits that the infectious agent is composed solely of the misfolded conformer PrPSc. There have been many attempts to create infectious prions from purified recombinant PrP protein. However, all of the samples generated in these experiments display relatively low levels of specific infectivity when inoculated intracerebrally into wild-type animals.14 Several lines of evidence suggest that cellular cofactors, such as polyanionic molecules, facilitate the formation of the infectious conformation.514The first in vitro PrP conversion assay used radiolabeled PrPC substrate purified from mammalian cells mixed with a stoichiometric excess of unlabeled PrPSc. This cell free assay produced a protease-resistant, radioactive product termed PrP-res.15 These pioneering studies showed for the first time that PrP could be specifically transformed in vitro, but the yield using purified substrates was low. Using a modification of the cell free assay in which crude brain homogenate replaced purified PrPC as the substrate, our laboratory was able to amplify PrPSc 6-fold over input prion seed, suggesting that non-PrP constituents of crude brain homogenate might be required for efficient PrPSc formation in vitro.16 Using this system, we discovered that nuclease treatment of hamster brain homogenates abolished PrPSc amplification in vitro, and that reconstituting the nuclease-treated reactions with purified mammalian RNA rescued the amplification process.5 PrPSc amplification could also be obtained by adding certain synthetic homopolymeric nucleic acids to immunopurified PrPC.6 Taken together, these surprising results argue that non-proteinaceous, host-encoded cofactors such as RNA molecules might facilitate prion conversion through a structural (as opposed to encoding) mechanism.8 The high efficiency of the serial protein misfolding amplification (sPMCA) technique developed by Soto and colleagues has allowed researchers to amplify prion infectivity as well as PrPSc molecules.17,18 Using sPMCA, we showed that infectious PrPSc molecules could be formed from immunopurified PrPC, co-purified lipid and synthetic RNA molecules. Moreover, even unseeded reactions containing these defined components were capable of generating prions with high specific infectivity in a prion-free environment, showing for the first time that wild type infectious prions could be produced de novo.7Additional studies in this purified system showed that PrPC molecules undergo a time-dependent conformational change upon interaction with RNA. When this change occurs, PrPC adopts an intermediate conformation that mimics some of the characteristics of PrPSc, such as detergent insolubility and reactivity to PrPSc-specific antibodies, but remains sensitive to proteinase K digestion.8 When incubated with a heterogeneous size mixture of homopolymeric [32P] poly(A) molecules during PMCA, hamster PrPC molecules incorporated a specific size subset (1–2.5 kb) of the RNA molecules into nuclease-resistant complexes. The physical interaction between RNA and PrPSc was confirmed by fluorescence microscopy experiments showing that fluorescein-labeled RNA molecules became integrated into nuclease-resistant complexes with PrPSc molecules. Interestingly, neuropathologic analysis of scrapie-infected hamsters revealed that endogenous RNA molecules stained with acridine orange co-localized with large extracellular PrP aggregates.8 Taken together, these studies suggest that PrP interacts specifically with polyanionic molecules in vitro and in situ, and raised the possibility that polyanions might be a necessary component of infectious prions.Jeong et al. investigated whether endogenous RNA molecules might be required for prion infectivity by treating scrapie brain homogenates with LiAlH4 (lithium aluminum hydride), a strong reducing agent that can cleave the phosphodiester bond in RNA molecules.19 Interestingly, treatment of hamster scrapie brain homogenates with LiAlH4 caused an ∼3-fold increase in scrapie incubation period measured by bioassay, suggesting that RNA may be an important component of infectious prions and therefore may play a role in stabilizing PrPSc structure. However, LiAlH4 is not a specific reagent, and can damage a variety of other macromolecules, including proteins. Therefore, the decrease in infectivity measured in this study cannot be specifically ascribed to degradation of the polyanion.19We recently reinvestigated the potential role of polyanion in maintaining prion infectivity by using a more targeted approach.20 Specifically, we utilized a synthetic oligonucleotide that could be selectively hydrolyzed by treatment with ultraviolet (UV) light. The photocleavable oligonucleotide was synthesized by inserting a photocleavable linker in between every fives bases of a poly(dT) 100-mer. Exposure to UV light quantitatively converted the oligonucleotide into five base fragments. During incubation with excess recombinant PrP, the photocleavable oligonucleotide became incorporated into a nuclease-resistant nucleoprotein complex, but remained sensitive to photocleavage. This novel system allowed us to study the role of a polyanion molecule incorporated into infectious prions in situ (Fig. 1).Open in a separate windowFigure 1Selective photodegradation of an incorporated polyanion in situ.We used PMCA to create PrPSc molecules that contained either the photocleavable oligonucleotide or a non-photocleavable control analog. After treatment with UV light, the infectivity of each sample was measured using a combination of end-point dilution sPMCA and animal bioassays. The end-point dilution PMCA assay showed a ∼1 log decrease in the seeding ability of PrPSc samples treated with UV light, but this effect was not specific since a similar decrease was measured in samples containing the control nucleic acid. In the bioassay, there was no change in the incubation periods of animals inoculated with PrPSc samples treated either in the presence or absence of UV light. Neuropathological analysis of inoculated animals also showed no differences in neurotropism between the two groups. Degradation of the nucleic acid had no effect on the molecular migration or structural stability of PrPSc samples as determined by SDS-PAGE and urea denaturation assays, respectively. There were also no differences in the molecular migration or glycosylation profile of the PrPSc molecules produced in the brains of animals inoculated with light- versus mock-treated inocula, and urea denaturation assays showed no differences in PrPSc stability. These results collectively demonstrate that the presence of intact polyanion molecules is not required to maintain the infectious, biochemical or strain properties prions generated in vitro.These results are consistent with the stringent “protein-only” hypothesis, but do not yet provide definitive proof. The purified PrPC molecules used as substrate in these experiments contain a stoichiometric amount of co-purified lipid7 that may play a role in the generation of prion infectivity.9 Also, although the efficacy of photocleavage conditions was carefully confirmed in control reactions, it is possible that some intact oligonucleotide survived UV treatment at a level below detection. Alternatively, the remnant five base nucleic acid fragments may remain incorporated within the PrPSc molecule and play a role in maintaining the infectious conformation. Even in this scenario, our results would place a significant geometric constraint on the role of incorporated polyanion. While polyanions ≥40 bases facilitate the formation infectious prions in vitro,8 our results suggest that polyanions >5 bases are not necessary to maintain the infectious properties the prion. The exact role polyanions play in prion formation is still unclear, but it is tempting to speculate that they may serve as scaffolds that facilitate prion conversion by (a) bringing PrPC and PrPSc seed together for templating to occur or (b) acting as a catalyst which is necessary to reduce the activation energy of refolding to the PrPSc form. Future studies will need to be performed to differentiate between these two hypotheses. It is also possible that polyanions are completely dispensable for maintaining PrPSc structure, and it is the co-purified lipid molecules that serve this role instead. Consistent with this possibility, we recently discovered that mouse PrPSc can be serially propagated in vitro in the absence of nucleic acids.21 Finally, it is possible that either polyanions or lipids can function equally well as stabilizers of the infectious PrPSc conformation. More work is required to distinguish between these possibilities.Generating high levels of specific infectivity solely using purified recombinant PrP remains the ultimate proof of the “protein-only” hypothesis. To date, evidence suggests that cellular cofactors are necessary to create infectious prions but may or may not be required to maintain infectivity once formed. Significantly, Wang et al. showed that bona fide prions could be formed from recombinant PrP, synthetic lipid and RNA molecules.9 Although no completely pure preparations of misfolded PrP possessing significant levels of specific infectivity have yet been produced, it should eventually be possible to produce such a preparation if the “protein-only” hypothesis is correct. On the other hand, a rigorous refutation of the hypothesis would require demonstrating that PrPSc and infectivity can be dissociated.  相似文献   

2.
3.
Neurodegenerative diseases are caused by proteinaceous aggregates, usually consisting of misfolded proteins which are often typified by a high proportion of β-sheets that accumulate in the central nervous system. These diseases, including Morbus Alzheimer, Parkinson disease and Transmissible Spongiform Encephalopathies (TSEs)—also termed prion disorders—afflict a substantial proportion of the human population and, as such, the etiology and pathogenesis of these diseases has been the focus of mounting research. Although many of these diseases arise from genetic mutations or are sporadic in nature, the possible horizontal transmissibility of neurodegenerative diseases poses a great threat to population health. In this article we discuss recent studies that suggest that the “non-transmissible” status bestowed upon Alzheimer and Parkinson diseases may need to be revised as these diseases have been successfully induced through tissue transplants. Furthermore, we highlight the importance of investigating the “natural” mechanism of prion transmission including peroral and perenteral transmission, proposed routes of gastrointestinal uptake and neuroinvasion of ingested infectious prion proteins. We examine the multitude of factors which may influence oral transmissibility and discuss the zoonotic threats that Chronic Wasting disease (CWD), Bovine Spongiform Encephalopathy (BSE) and Scrapie may pose resulting in vCJD or related disorders. In addition, we suggest that the 37 kDa/67 kDa laminin receptor on the cell surface of enterocytes, a major cell population in the intestine, may play an important role in the intestinal pathophysiology of alimentary prion infections.Key words: prion, 37 kDa/67 kDa laminin receptor, CJD, BSE, CWD, scrapie, Alzheimer disease, Parkinson disease, intestine, enterocytesMany different mechanisms exist which underlie the etiology of the numerous neurodegenerative diseases affecting the human population. Amongst the most prominent are Morbus Alzheimer, prion disorders, Parkinson disease, Chorea Huntington, frontotemporal dementia and amylotrophic lateral sclerosis. The molecular mechanisms underlying these diseases vary; however, all neurodegenerative diseases share a common feature: they are caused by protein aggregation. The only neurodegenerative diseases proven to be transmissible are prion disorders. In contrast to frontotemporal dementia, recent evidence suggests that Alzheimer and Parkinson diseases may also be transmissible. Pre-symptomatic Alzheimer disease (APP23) mice exhibited an increase in the Alzheimer phenotype when brain homogenate of autopsied human Alzheimer disease patients and older, amyloid beta- (Aβ-) laden APP23 mice was injected into their hippocampi.1 These findings suggest that the Aβ-abundant brain homogenate of Alzheimer disease patients may possess the ability to induce or supplement the overproduction of Aβ, possibly leading to the onset of Alzheimer disease.The pathological feature associated with Parkinson disease is the formation of Lewy bodies in cell bodies and neuronal processes in the brain.2 The main component of these protein aggregates is α-synuclein (reviewed in ref. 2). Autopsies of Parkinson disease patients revealed that Lewy bodies had formed on healthy embryonic neurons that had been grafted onto the brain tissue of the patients several years before (prior to said examination).35 It may thus be proposed that α-synuclein transmission is possible from diseased to healthy neurons, suggesting that Parkinson disease may be transmissible from a Parkinson disease patient to a healthy individual. These findings imply that Alzheimer and Parkinson diseases may be transmissible through tissue transplants and the use of contaminated surgical tools.6Prion disorders, also termed Transmissible Spongiform Encephalopathies (TSEs), are fatal neurodegenerative diseases that affect the central nervous system (CNS) of multiple animal species. In lieu of the social, economic and political ramifications of such infections, as well as the possible intra- and interspecies transmissibility of such disorders, various routes of experimental transmission have been investigated including intracerebral, intraperitoneal, intraventricular, intraocular, intraspinal and subcutaneous injections (reviewed in ref. 79). However, such routes of transmission are not representative of the “natural” mechanism as the majority of prion disorders are contracted through ingestion of infectious prion (PrPSc) containing material. Thus, the peroral and perenteral prion transmission is of greatest consequence with respect to TSE disease establishment. Moreover, the presence of PrPSc in the buccal cavity of scrapie-infected sheep10 (reviewed in ref. 11) and the possible horizontal transfer as a result hereof, as may be similarly proposed for animals suffering from other TSEs, may further contribute to the oral transmissibility of TSEs.A number of model systems have been employed to study TSE transmissibility. Owing to ethical constraints, TSE transmissibility to humans via the oral route may not be directly investigated and as a result hereof, alternative model systems are needed. These may include the use of transgenic mice, cell lines which are permissive to infection12 and experimental animals such as sheep, calves, goats, minks, ferrets and non-human primates (reviewed in ref. 9).Intestinal entry of PrPSc has been proposed to occur via two pathways, the membranous (M) cell-dependent and M cell-independent pathways (Fig. 1).13,14 The former involves endocytic M (microfold)-cells, which cover the intestinal lymphoid follicles (Peyer''s patches)14 and may take up prions and thereby facilitate the translocation of these proteins across the intestinal epithelium into the lymphoid tissues (reviewed in ref. 9) as has been demonstrated in a cellular model.13 Following such uptake by the M cells, the prions may subsequently pass to the dendritic cells and follicular dendritic cells (FDCs) (Fig. 1), which allow for prion transport to the mesenteric lymph nodes and replication, respectively.15 The prion proteins may subsequently gain access to the enteric nervous system (ENS) and ultimately the central nervous system (CNS).15Open in a separate windowFigure 1Proposed routes of gastrointestinal entry of ingested infectious prions (PrPSc) as well as possible pathways of amplification and transport to the central nervous system.However, prion intestinal translocation has been observed in the absence of M cells and has been demonstrated to be as a result of the action of polar, 37 kDa/67 kDa LRP/LR (non-integrin laminin receptor; reviewed in ref. 1618) expressing enterocytes. Enterocytes are the major cell population of the intestinal epithelium and due to their ability to endocytose pathogens, nutrients and macromolecules,19 it has been proposed that these cells may represent a major entry site for alimentary prions (Fig. 1).Since enterocyte prion uptake has been demonstrated to be dependent on the presence of LRP/LR on the apical brush border of the cells,14,20 the interaction between varying prion protein strains and the receptor2123 may be employed as a model system to study possible oral transmissibility of prion disorders across species as well as the intestinal pathophysiology of alimentary prion infections.24 Moreover, the blockage of such interactions through the use of anti-LRP/LR specific antibodies has been reported to reduce PrPSc endocytosis19 and thus these antibodies may serve as potential therapeutics to prevent infectious prion internalization and thereby prevent prion infections. It must be emphasized that the adhesion of prion proteins to cells is not solely dependent on the LRP/LR-PrPSc interactions;24 however, this interaction is of importance with regards to internalization and subsequent pathogenesis.We applied the aforementioned cell model to study the possible oral transmission of PrPBSE, PrPCWD and ovine PrPSc to cervids, cattle, swine and humans.24 The direct transmission of the aforementioned animal prion disorders to humans as a result of dietary exposure and the possible establishment of zoonotic diseases is of great public concern. It must however be emphasized that the study investigated the co-localization of LRP/LR and various prion strains and not the actual internalization process.PrPBSE was shown to co-localize with LRP/LR on human enterocytes24, thereby suggesting that PrPBSE is transmissible to humans via the oral route which is widely accepted as the manner by which variant CJD originated. This suspicion was previously investigated using a macaque model, which was successfully perorally infected by BSE-contaminated material and subsequently lead to the development of a prion disorder that resembles vCJD.25 These results, due to the evolutionary relatedness between macaques and humans, allowed researchers to confirm the oral transmissibility of PrPBSE to humans. PrPBSE may also potentially lead to prion disorder establishment in swine,24 livestock of great economic and social importance.The prion disorder affecting elk, mule deer and white-tailed deer is termed CWD. Cases of the disease are most prevalent in the US but are also evident in Canada and South Korea.26,27 As the infectious prion isoform is reported to be present in the blood28 and skeletal muscle,29 hunting, consumption of wild venison and contact with other animal products derived from CWD-infected elk and deer may thereby pose a public health risk. Our studies demonstrate that PrPCWD co-localizes with LRP/LR on human enterocytes24 thereby suggesting a possible oral transmissibilty of this TSE to humans. This is, however, inconsistent with results obtained during intra-cerebral inoculation of the brains and spinal cords of transgenic mice overexpressing the human cellular prion protein (PrPc),26,27 which is essential for TSE disease establishment and progression. Further, discrepancies have also been reported with respect to non-human primates, as squirrel monkeys have been successfully intracerebrally inoculated with mule-deer prion homogenates,30 while cynolmolgus macaques were resistant to infection.31 CWD has been transmitted to ferrets, minks and goats32 and as these animals may serve as domestic animals or livestock, secondary transmission from such animals to humans, through direct contact or ingestion of infected material, may be an additional risk factor that merits further scientific investigation.Ovine PrPSc co-localization with LRP/LR on human and bovine enterocytes may be indicative of the infectious agents'' ability to effect cross-species infections. The oral transmissibility of Scrapie has been confirmed in hamsters fed with sheep-scrapie-infected material.33The discrepancies with regards to the transmissibility of certain infectious prion proteins when assessed by different model systems may be due to the experimental transmission route employed. Oral exposure often results in significantly prolonged incubation times when compared to intracerebral inoculation techniques and thus failure of transgenic mice and normal experimental animals to develop disease phenotypes after being fed TSE-contaminated material may not necessarily indicate that the infection process failed.14 Apart from the route of infection, numerous other factors may influence transmission between species, including dose, PrP polymorphisms and genetic factors, the prion strain employed as well as the efficacy of prion transport to the CNS.34 The degree of homology between the PrPc protein in the animals serving as the infectious prion source and recipient has also been described as a feature limiting cross-species transmission.34 The negative results, as referred to above, obtained upon prion-protein inoculation of animal models may have resulted due to the slow rate at which the infectious prion induces conformational conversion of the endogenous PrPc in the animal cells and this in turn results in low levels of infectious prion replication and symptom development.27Furthermore, even in the event that certain prion disorders are not directly transmissible to humans, most are transmissible to at least a single species of domestic animal or livestock. The infectious agents properties may be altered in the secondary host such that it becomes transmissible to humans (reviewed in ref. 35). Thus, interspecies transmission between animals may indirectly influence human health.It is noteworthy to add that although the oral route of PrPSc transmission may result in prolonged incubation times, it may broaden the range of susceptible hosts. A common constituent of food is ferritin, a protein that is resistant to digestive enzyme hydrolysis and, due to its homology across species, it may serve as co-transporter of PrPSc and facilitate enterocyte internalization of the infectious prion.36 It may thus be proposed that prion internalization may occur via a ferritin-PrPSc complex even in the absence of co-localization between the infectious agent and LRP/LR such that many more cross-species infections (provided that the other infection factors are favorable) may be probable.37 In addition, digestive enzymes in the gastrointestinal tract facilitate PrPSc binding to the intestinal epithelium and subsequent intestinal uptake36 and thus depending on the individuals'' digestive processes, the susceptibility to infection and the rate of disease development may vary accordingly. As a result hereof, though laboratory experiments in cell-culture and animal models may render a particular prion disorder non-infectious to humans, this may not be true for all individuals.In lieu of the above statements, with particular reference to inconsistencies in reported results and the multiple factors influencing oral transmissibility of TSEs, further transmission studies are required to evaluate the zoonotic threat which CWD, BSE and Scrapie may pose through ingestion.  相似文献   

4.
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   

5.
6.
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   

7.
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   

8.
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   

9.
The detailed structures of prion disease-associated, partially protease-resistant forms of prion protein (e.g. PrPSc) are largely unknown. PrPSc appears to propagate itself by autocatalyzing the conformational conversion and oligomerization of normal prion protein (PrPC). One manifestation of PrPSc templating activity is its ability, in protein misfolding cyclic amplification reactions, to seed the conversion of recombinant prion protein (rPrP) into aggregates that more closely resemble PrPSc than spontaneously nucleated rPrP amyloids in terms of proteolytic fragmentation and infrared spectra. The absence of posttranslational modifications makes these rPrP aggregates more amenable to detailed structural analyses than bona fide PrPSc. Here, we compare the structures of PrPSc-seeded and spontaneously nucleated aggregates of hamster rPrP by using H/D exchange coupled with mass spectrometry. In spontaneously formed fibrils, very slow H/D exchange in region ∼163–223 represents a systematically H-bonded cross-β amyloid core structure. PrPSc-seeded aggregates have a subpopulation of molecules in which this core region extends N-terminally as far as to residue ∼145, and there is a significant degree of order within residues ∼117–133. The formation of tightly H-bonded structures by these more N-terminal residues may account partially for the generation of longer protease-resistant regions in the PrPSc-seeded rPrP aggregates; however, part of the added protease resistance is dependent on the presence of SDS during proteolysis, emphasizing the multifactorial influences on proteolytic fragmentation patterns. These results demonstrate that PrPSc has a distinct templating activity that induces ordered, systematically H-bonded structure in regions that are dynamic and poorly defined in spontaneously formed aggregates of rPrP.Transmissible spongiform encephalopathies (TSEs),2 or prion diseases, are a group of infectious neurodegenerative disorders that affect many mammalian species and include Creutzfeldt-Jakob disease in humans, scrapie in sheep, chronic wasting disease in cervids, and bovine spongiform encephalopathy (“mad cow” disease) (17). All of these diseases appear to be intimately associated with conformational conversion of the normal host-encoded prion protein, termed PrPC, to a pathological isoform, PrPSc (15). According to the “protein-only” model, PrPSc itself represents the infectious prion agent (1, 8); it is believed to self-propagate by an autocatalytic mechanism involving binding to PrPC and templating the conversion of the latter protein to the PrPSc state (9, 10). Although molecular details of such a mechanism of disease propagation remain largely unknown, the general principle of protein-based infectivity is supported by a wealth of experimental data (17).PrPC is a monomeric glycophosphatidylinositol-linked glycoprotein that is highly protease-sensitive and soluble in nonionic detergents. High resolution NMR data show that the recombinant PrP (rPrP), a nonglycosylated model of PrPC, consists of a flexible N-terminal region and a folded C-terminal domain encompassing three α-helices and two short β-strands (1113). Conversely, the PrPSc isoform is aggregate in nature, rich in β-sheet structure, insoluble in nonionic detergents, and partially resistant to proteinase K (PK) digestion, with a PK-resistant core encompassing the C-terminal ∼140 residues (15, 14, 15). Little specific structural information is available, however, for this isoform beyond low resolution biochemical and spectroscopic characterization. Thus, the structure of PrPSc conformer(s) associated with prion infectivity remains one of the best guarded mysteries, hindering efforts to understand the molecular basis of TSE diseases.Many efforts have been made over the years to recapitulate PrPSc formation and prion propagation in vitro. Early studies have shown that PrPC can be converted with remarkable species and strain specificities to a PrPSc-like conformation (as judged by PK resistance) simply by incubation with PrPSc from prion-infected animals (16, 17). The yields of these original cell-free conversion experiments were low, and no new infectivity could be attributed to the newly converted material (18). An important more recent study showed that both PrPSc and TSE infectivity can be amplified indefinitely in crude brain homogenates using successive rounds of sonication and incubation (19), a procedure called protein misfolding cyclic amplification (PMCA) (20). Similar amplification of the TSE infectivity was also accomplished by PMCA employing purified PrPC as a substrate, although only in the presence of polyanions such as RNA and copurified lipids (21). Unfortunately, the quantities of infectious PrPSc generated by PMCA using purified brain-derived PrPC are very small, precluding most structural studies.In contrast to brain-derived PrPC, large scale purification can be readily accomplished for bacterially expressed rPrP, a form of PrP lacking glycosylation and the glycophosphatidylinositol anchor. The latter protein can spontaneously polymerize into amyloid fibrils, and much insight has been gained into mechanistic and structural aspects of this reaction (2228). However, although rPrP fibrils were shown to cause or accelerate a transmissible neurodegenerative disorder in transgenic mice overexpressing a PrPC variant encompassing residues 89–231, the infectivity titer of these “synthetic prions” was extremely low (29) or absent altogether (4). This low infectivity coincides with much shorter PK-resistant core of rPrP amyloid fibrils compared with brain-derived PrPSc (26, 30), raising questions regarding the relationship between these fibrils and the authentic TSE agent. In this context, an important recent development was the finding that the PrPSc-seeded PMCA method can be extended to rPrP, yielding protease-resistant recombinant PrP aggregates (rPrPPMCA or rPrP-res(Sc)) (31). These aggregates display a PK digestion pattern that is much more closely related to PrPSc than that of previously studied spontaneously formed rPrP fibrils, offering a potentially more relevant model for biochemical and biophysical studies. Here, we provide, for the first time, a direct insight into the structure of rPrPPMCA. H/D exchange data coupled with MS analysis (HXMS) allowed us to identify systematically H-bonded core region(s) of these aggregates, shedding a new light on the mechanisms underlying formation of PK-resistant structures.  相似文献   

10.
11.
The discovery of tunnelling nanotubes (TNTs) and their proposed role in long intercellular transport of organelles, bacteria and viruses have led us to examine their potential role during prion spreading. We have recently shown that these membrane bridges can form between neuronal cells, as well as between dendritic cells and primary neurons and that both endogenous and exogenous PrPSc appear to traffic through these structures between infected and non-infected cells. Furthermore, prion infection can be efficiently transmitted from infected dendritic cells to primary neurons only in co-culture conditions permissive for TNT formation. Therefore, we propose a role for TNTs during prion spreading from the periphery to the central nervous system (CNS). Here, we discuss some of the key steps where TNTs might play a role during prion neuroinvasion.Key words: tunnelling nanotubes, TNTs, prion, PrPSc, prion spreading, dendritic cells, neuroinvasionPrion diseases, or transmissible spongiform encephalopathies (TSEs), are fatal neurodegenerative disorders that have been found in a number of species, including scrapie in sheep, bovine spongiform encephalopathy in cattle (BSE), Chronic wasting disease in deer and Creutzfeldt-Jacob, the Gerstmann-Straüssler-Scheinker syndrome, fatal familial insomnia and kuru in humans (reviewed in ref. 1). Human TSEs can be sporadic, genetic or acquired by infection. A new variant of Creutzfeldt-Jakob disease (termed vCJD) was reported from the UK in 1996.2 The majority of vCJD cases diagnosed to date resulted from a peripheral exposure via the consumption of BSE-contaminated food. Pathological features of TSE diseases can include gliosis, neuronal cell loss and spongiform changes, but the common feature of all members of this group of diseases is the build-up of an aberrant form of the host cellular protein PrPC, named PrPSc (from scrapie). The normal cellular isoform, PrPC, is an endogenous glycosylphosphatidyl inositol (GPI)-anchored protein present in numerous tissues in mammals, including neurons and lymphoid cells. While the exact function of PrPC remains unclear, evidence suggest putative roles in neuroprotection, cell adhesion and signal transduction (reviewed in refs. 3 and 4). According to the ‘protein-only hypothesis,’ the causative agents of prion diseases are proteinaceous infectious particles (‘prions’), which are composed essentially of misfolded PrPC, or PrPSc.5,6 Prions replicate through a molecular mechanism in which abnormally folded PrPSc acts as a catalyst and serves as a template to convert normal PrPC molecules into PrPSc.5,6 PrPSc differs from PrPC in the conformation of its polypeptide chain, which is enriched in β-sheets and is protease resistant. Although the conversion process is believed to have a predominant role in the pathogenesis of prion diseases, the cellular and molecular basis for the pathogenic conversion of PrP are still unknown.Another important question is how PrPSc spreads to and within the brain. After oral exposure, PrPSc accumulates into lymphoid tissues, such as the spleen, lymph nodes or Peyer''s patches, prior to neuroinvasion.79 The exact mechanisms and specific cells involved in the spreading from the gastrointestinal track to the lymphoid system and to the peripheral nervous system (PNS), leading to neuroinvasion of the CNS remain to be elucidated. However, a range of evidence suggests that the accumulation of PrPSc within lymphoid tissues is necessary for efficient neuroinvasion.911 In particular it has been shown that PrPSc accumulates first within follicular dendritic cells (FDCs)12 and macrophages.13 FDCs are stromal-differentiated cells in the germinal centres of activated lymphoid follicles. A number of studies have demonstrated that FDCs play a critical role during spreading of infection since their absence greatly impaires the neuroinvasion process.8,11,14,15 However, because FDCs are immobile cells, it is not clear how they acquire PrPSc and how it spreads from the FDCs to the PNS. FDCs and nerve synapses occupy different anatomical sites16,17 and therefore the lack of physical contact between the gut and FDCs and between FDCs and the nerve periphery imply the presence of intermediate mechanisms for the transport of PrPSc. Dendritic cells (DCs) have been proposed to play a critical role in the transport of PrPSc from the gut to FDCs.18 DCs function as sentinels for incoming pathogens. Bone-marrow dendritic cells (BMDCs) are migratory cells that are able to transport proteins within Peyer''s patches and into mesenteric lymph nodes.19 Interestingly, mucosal dendritic cells which play a role in the transport of intestinal antigen for presentation to Peyer''s patches and to mesenteric lymph nodes, can also extend trans-epithelial dendrites to directly sample bacteria in the gut.20,21 However, the transport of PrPSc from FDCs to the PNS remains controversial and evidence for a direct role of DCs during this process has been debated.22,23 Several mechanisms have been proposed for the intercellular transfer of PrPSc, including cell-cell contact, transfer via exosomes or by GPI-painting.2426 For example, similar to other types of pathogens such as HIV-1, which was proposed to follow the “exosomal” pathway to be released from the cells,27 it has been shown that the supernatant of prion infected cells contain large amount of PrPSc in membranous vesicles known as exosomes.25,28 Thus, it was suggested that exosomes might be a way to spread prion infection in vivo.25,28 Recently, a different type of vesicles known as plasma membrane-derived microvesicles, were also described as a potential spreading mechanism during neuroinvasion.29In 2004, Rustom and colleagues discovered a new mechanism of long distance intercellular communication in mammalian cells, called tunnelling nanotubes (TNTs).30 TNTs are transient, long, actin-rich projections that allow for long-distance intercellular communication (reviewed in refs. 3133). TNT-like structures have been described to form in vitro between numerous cell types, including neuronal and immune cells.30,34,35 These studies demonstrated that TNT-like structures formed bridges or channels between distant cells that can be used to transfer material between cells, including Lysotracker positive or endosomal vesicles, calcium fluxes, bacteria or viruses through their cytoplasms or along the surface of the nanotubes.3133 Interestingly, a model GPI-anchored protein, GFP-GPI, was found to move at the surface of these tubes34 and while studying the neuritic transport of prions in neuronal cells, Magalhães and colleagues noticed a strong correlation between internalized PrP-res and Lysotracker positive vesicles in neurites,36 suggesting that PrP-res might also be able to transfer through TNTs during prion cell-cell spreading.The results from the studies mentioned above and random observations of TNT-like structures in neuronal model cell cultures first led us to study whether these structures could in fact provide an efficient mechanism for prion cell-cell spreading.37 We initially characterized TNT-like structures in the mouse catecholaminergic neuronal cell line, Cath.a-Differentiated cells (CAD cells) a well-recognized neuronal cell model for prion infection.38 Under our culturing conditions, over 40% of the CAD cells could efficiently form actin-rich TNT-like structures between differentially labelled cell populations. In CAD cells, these nanotubes were very heterogeneous, both in length and in diameters. Indeed, TNT-like structures had lengths ranging from 10 to 80 µm and while over 70% of the nanotubes had diameters smaller than 200 nm, the remaining TNT-like structures had larger diameters (200 to 800 nm). We demonstrated that vesicles of lysosomal origins, a fluorescent form of PrP (GFP-PrP), infectious Alexa-PrPSc, as well as both endogenous and exogenous PrPSc could traffic within TNTs between neuronal cells (Fig. 1). The lysosomal and GFP-PrP vesicles observed to move through TNTs had a directed movement with a speed in the range of actin-mediated motors,37 consistent with previous studies suggesting the involvement of an actomyosin-dependent transport.39 Interestingly, active transfer of endogenous PrPSc, lysosomal or GFP-PrP vesicles occurred through TNTs with larger diameters, suggesting distinct roles for the different TNT-like structures observed.37 These results do not seem to be specific to CAD cells since the transfer of GFP-PrP throught TNTs was observed in different types of transfected cells, including HEK293 cells (unpublished data). Furthermore, these results were in agreement with previous observations by Onfelt and colleagues showing the presence of a fluorescent GPI model protein (GFP-GPI) in TNTs formed between EBV-transformed human B cells34 suggesting that different GPI-anchored proteins can be transferred along the surface and inside vesicles within TNTs. In order to determine the relevance of this type of intercellular communication in the case of prion diseases, it was necessary to evaluate the trafficking of the pathological form of PrP (PrPSc) within TNTs, by analyzing the transfer of endogenous PrPSc between chronically infected ScCAD cells and non-infected CAD cells. By immunofluorescence after guanidium treatment, endogenous PrPSc was found inside TNTs and in the cytoplasm of recipient non-infected CAD cells. Similar to exogenous PrPSc, endogenous PrPSc particles were not present in non-infected CAD cells not in contact with ScCAD cells after overnight co-cultures, thus excluding exosomal transfer or protein shedding.37 Similarly, no transfer was observed between cells in direct contact with one another or upon treatment with latrunculin, which inhibits TNT formation. Strikingly, the transfer of endogenous PrPSc was visible only when TNTs were present, demonstrating that in vitro, PrPSc can efficiently exploit TNTs to spread between cells of neuronal origin. These data suggested that TNTs could be a mechanism for prion spreading within the cells of the CNS.Open in a separate windowFigure 1Endogenous PrPSc transfer from ScCAD cells to CAD cells via TNTs. Endogenous PrPSc is found in punctate structures inside TNTs and in the cytoplasms of recipient cells. CAD cells were transfected with Cherry-PLAP (red) and co-cultured with ScCAD for 24 h. Cells were fixed, treated with Gnd and immunostained for PrP using SAF32 Ab (green). (A) Merge projection of Z-stacks obtained with a confocal Andor spinning-disk microscope. (B) Three-dimensional reconstruction of (A) using OsiriX software. (C) Zoom in on TNT-like structures. PrPSc is found in vesicular structures inside TNTs and in the cytoplasm of the recipient non-infected CAD cells (see blue arrow heads). Scale bar represents 10 µm.Interestingly, DCs were shown to form networks of TNTs both in vitro40 and in vivo.41 In an elegant study, Watkins and Salter demonstrated that DCs could propagate calcium flux upon cell stimulation to other cells hundreds of microns away through TNTs, both between DCs and between DCs and THP-1 monocytes.40 These data suggested the possibility that DCs could form tubular connections with neuronal cells in order to transport PrPSc to the PNS via TNTs. Using BMDCs in co-cultures with both cerebellar granular neurons (CGNs) and primary hippocampal neurons, we showed that BMDCs could form networks of TNTs with both types of neurons. Furthermore, these TNTs appeared to be functional, allowing for the transport of Lysotracker positive vesicles and infectious Alexa-PrPSc between loaded BMDCs and primary neurons, suggesting that DCs could transfer the infectious prion agent to primary neuronal cultures through TNTs. By using filters and conditions unfavorable for other mechanisms of transport, we found that moRK13 cells,28 as well as CGNs (unpublished data), could be infected by co-cultures with BMDCs loaded with infectious brain homogenate.37 Overall, these data indicate that TNTs could be an efficient mechanism of prion transmission between immune cells and neuronal cells, as well as between neuronal cultures. Since DCs can interact with peripheral neurons,42 we propose that TNTs could be involved in the process of neuroinvasion at multiple stages, from the peripheral site of entry to the PNS by neuroimmune interactions with DCs, allowing neurons to retrogradely transport prions to the CNS, and within the CNS (Fig. 2).Open in a separate windowFigure 2Transport of PrPSc via TNTs, an alternative spreading mechanism during neuroinvasion. Studies in our laboratory suggest that TNTs allow for the intracellular transport of PrPSc between dendritic cells and neurons and between neurons (see inset). The exact mechanism of transport remains to be determined. For instance, it is still not clear, whether PrPSc is strictly transported within endocytic vesicles, or whether it can slide along the surface or be transported as aggregosomes within the tubes. Similarly, the types of motors used, as well as the possible gated mechanisms to enter the recipient cells are not known. Because of the high propensity of DCs to form TNTs with different cell types, we propose that TNTs could play important roles in delivering PrPSc to the proper cell types along the neuroinvasion route. For instance, DCs could deliver PrPSc from the peripheral entry sites to FDCs in the secondary lymphoid tissues (2) or in a less efficient manner, they might occasionally directly transport PrPSc to the PNS (1). They could also bridge the immobile FDC networks and the PNS (3), since we have shown that DCs can form TNTs with nerve cells. Finally, once PrPSc has reached its final destination within the CNS, TNTs might play a final role in the spreading of PrPSc within the brain between neurons and possibly between neuronal cells and astrocytes (4).Recently, it was demonstrated that the distance between FDCs and the neighbouring PNS was critical for prion neuroinvasion.43 Indeed, in the spleen of CD19−/− mice, FDC networks were found to be 50% closer to the nerve fibers compared to wild-type mice.43 The authors suggested that the increase in prion spreading efficiency in these mice was directly dependent on the reduction in the distance between the FDC networks and the PNS in these mice. These results would be consistent with a mechanism of transfer such as exosomes release. However, shortening the distance between FDCs and the PNS would also reduce the route of transport that mobile cells would have to travel and increase the chances for transfer of prions to the PNS, resulting in an increase in prion spreading efficiency. While the importance of FDCs in prion replication during the spreading to the CNS seems to be clear,11,14,15 their specific role in the transfer of prions and their possible interactions with other mobile cells are much more debated.22,23 In order to bridge the gap between FDCs and the PNS, a role for DCs as possible carriers of PrPSc has been postulated. Aucouturier and colleagues have previously shown, using RAG-1−/− mice, which are deficient in FDCs, that CD11c+ DCs infected with 139A were able to carry prion infection to the CNS, without accumulation and replication in lymphoid organs,22 thus suggesting that DCs are able to transport prions from the periphery to nerve cells. Recently, however, another study using TNFR1−/− mice, deficient for FDCs, suggested that DCs were unlikely candidates in the transport of prion to the PNS.23 In this study, the authors showed that in TNFR1−/− mice, ME7 or 139A infected DCs were inefficient in transferring infection to the PNS. The authors suggested that the differences between the results obtained with RAG-1−/− mice and TNFR1−/− mice could be due to the differences in the levels of innervation of the spleen in RAG-1−/−mice compared to TNFR1−/− mice. They suggested that in RAG-1−/− mice, DCs could spread prion infection because their spleens are highly innervated, compared to TNFR1−/− or wild-type mice, therefore increasing the propensity of DCs to encounter nerve cells and tranfer the prion agents. Because of the reduction in the levels of innervation in the spleens of wild-type mice, the authors concluded that DCs are unlikely candidate for the transport of prions directly to the PNS [see (1) in Fig. 2]. However, since these studies are using mice deficient for FDCs, it remains unclear what type of interactions might occur between FDCs and DCs, and how DCs might be able to transport prions from FDCs to the PNS in wild-type mice [see (2) in Fig. 2]. Indeed, both studies show that under the right circumstances, DCs can interact with nerve cells, similar to what was recently shown in infected mice42 and in agreement with our findings that DCs can form TNTs with neurons.37 Within this scenario, it is clear that to determine the specific role of DCs during the spreading of prions from the gut to the PNS, the transfer mechanisms between DCs and other cell types, especially FDCs and peripheral neurons, need to be better characterized.Overall, these in vitro data strongly point toward TNTs as one possible mechanism of prion spreading. The next step will be to identify these structures in vivo and to determine whether prion spreading in vivo is the result of passive mechanisms, such as exosome release, active intercellular transport along and within TNTs or whether prions will use any means available to reach their targets. Recently, TNT-like structures were imaged in a mouse cornea,41 suggesting that while challenging, the visualization of in vivo trafficking of prions in lymphoid tissues such as in lymph nodes or in the spleen as well as in brain organotypic cultures might be possible and could be used to reveal the presence of TNTs.The discovery of the existence of nanotubular membrane bridges in vitro has opened-up a new field of research. Channels, called plasmodesmata,44 connecting plant cells have long been known to play crucial roles in the transport of nutrients, molecules and signals during development and some of their functions were recently compared with some of the recently proposed functions of TNTs.45 Furthermore, in vivo long, actin-rich filopodia like structures were found to be crucial during development.4649 For example, these structures exist in developing sea urchin embryos and were proposed to play a role in signalling and patterning during gastrulation.47 Similar roles were proposed for thin filopodia-like structures observed during dorsal closure in drosophila.49 In addition, TNT-like structures were observed in the mouse cornea between DCs and were shown to increase under inflammatory conditions.41 The authors postulated that these TNT-like structures could play a role in Ag-specific signalling, especially as a response to eye inflammation. Therefore, the possibility that TNTs might play numerous roles during cell development, in the immune system and as conduits for the spreading of pathogens could lead to major changes in the way we view animal cell interactions. Specifically, understanding how pathogens usurp these cellular connections to spread could allow for the screening and the identification of new therapeutic inhibitors. To this aim, characterizing the basic mechanism of TNT formation within cell model systems will be necessary to improve the knowledge of TNTs in general, to analyze the transfer of pathogens more specifically, and to identify key molecules during this process. In the case of prions, whether they hijack nanotubes to spread between cells or whether prions increase the formation of filopodia and TNT-like structures similar to some viruses33,50 and/or the efficiency of transfer remain to be determined. Overall, in this specific field, the constant improvement of cell imaging techniques and the emergence of imaging tools to study prion spreading36,37,5153 could lead to exciting new insights both in the physiology of these intercellular connections and in the pathology of these devastating diseases.  相似文献   

12.
The cellular prion protein (PrPC) is essential for the pathogenesis and transmission of prion diseases. PrPC is bound to the plasma membrane via a glycosylphosphatidylinositol anchor, although a secreted, soluble form has also been identified. Previously we reported that PrPC is subject to ectodomain shedding from the membrane by zinc metalloproteinases with a similar inhibition profile to those involved in shedding the amyloid precursor protein. Here we have used gain-of-function (overexpression) and loss-of-function (small interfering RNA knockdown) experiments in cells to identify the ADAMs (a disintegrin and metalloproteinases) involved in the ectodomain shedding of PrPC. These experiments revealed that ADAM9 and ADAM10, but not ADAM17, are involved in the shedding of PrPC and that ADAM9 exerts its effect on PrPC shedding via ADAM10. Using dominant negative, catalytically inactive mutants, we show that the catalytic activity of ADAM9 is required for its effect on ADAM10. Mass spectrometric analysis revealed that ADAM10, but not ADAM9, cleaved PrP between Gly228 and Arg229, three residues away from the site of glycosylphosphatidylinositol anchor attachment. The shedding of another membrane protein, the amyloid precursor protein β-secretase BACE1, by ADAM9 is also mediated via ADAM10. Furthermore, we show that pharmacological inhibition of PrPC shedding or activation of both PrPC and PrPSc shedding by ADAM10 overexpression in scrapie-infected neuroblastoma N2a cells does not alter the formation of proteinase K-resistant PrPSc. Collectively, these data indicate that although PrPC can be shed through the action of ADAM family members, modulation of PrPC or PrPSc ectodomain shedding does not regulate prion conversion.The prion protein (PrP)3 is the causative agent of the transmissible spongiform encephalopathies such as Creutzfeldt-Jakob disease in humans, scrapie in sheep, bovine spongiform encephalopathy in cattle, and chronic wasting disease in deer and elk (1). In these diseases the cellular form of PrP (PrPC) undergoes a conformational conversion to the infectious form PrPSc that is characterized biochemically by its resistance to digestion with proteinase K (PK) (2). Mature PrPC is anchored to the extracellular surface of the cell membrane through a glycosylphosphatidylinositol (GPI) anchor and, like most GPI-anchored proteins, is clustered into cholesterol-rich, detergent-resistant membrane rafts (reviewed in Ref. 3). Although the precise subcellular site of conversion remains undefined, conformational conversion of PrPC to PrPSc is believed to occur either at the cell surface or within the endocytic pathway (46).A number of studies indicate that modulation of PrPC levels at the cell surface may represent a possible future disease intervention strategy. For example, the retention of PrPC at the cell surface and concomitant prevention of its endocytosis through the use of PrP antibodies inhibited PrPSc formation (7). Furthermore, the sterol-binding polyene antibiotic filipin reduced endocytosis, and induced cellular release, of PrPC with a concomitant reduction in PrPSc accumulation (8). More recently, it has been shown that modulation of cell surface PrPC levels by the novel sorting nexin SNX33 can interfere with PrPSc formation in cultured cells (9). Nonetheless, the natural processes regulating PrPC levels at the cell surface remain poorly defined. One such mechanism of regulation is via shedding of the bulk of the ectodomain of PrPC either through cleavage of the polypeptide close to the GPI anchor or within the GPI anchor itself. Indeed, it has long been established that PrPC can be shed into the medium of cultured cells and is present as a soluble form in vivo in human cerebrospinal fluid (10, 11).Numerous cell surface proteins can be proteolytically shed by the action of a group of zinc metalloproteinases known collectively as secretases or sheddases (reviewed in Refs. 12, 13). Whereas most proteolytically shed proteins are derived from transmembrane polypeptide-anchored substrates, several GPI-anchored proteins, including the folate receptor (14), the ecto-ADP-ribosyltransferase ART2.2 (15), and a GPI-anchored construct of angiotensin-converting enzyme (16) are shed by the action of metalloproteinases. We have previously shown that PrPC can also be proteolytically shed from the cell surface through the action of one or more zinc metalloproteinases with similar properties to those of the α-secretases responsible for the shedding of the amyloid precursor protein (APP) of Alzheimer disease (17). This α-secretase-mediated ectodomain shedding of APP from the cell surface is carried out by at least three members of the a disintegrin and metalloproteinase (ADAM) family, namely ADAM9, -10, and -17 (reviewed in Ref. 18). In addition to cleavage by ADAMs, APP is also cleaved by the β-secretase, BACE1 (β-site APP-cleaving enzyme) and the γ-secretase complex to release the neurotoxic amyloid-β peptide (19). BACE1 itself is also subject to ectodomain shedding by as yet unidentified members of the ADAM family (20).The similarities between the ectodomain shedding of APP and PrPC, in particular the similar profile of inhibition by a range of hydroxamate-based zinc metalloproteinase inhibitors (17), led us to investigate whether the same members of the ADAM family were also involved in the shedding of PrPC. It should be noted that this ectodomain shedding of PrPC by cleavage of the polypeptide chain near to the site (Ser231) of GPI anchor addition in the C terminus of the protein is distinct from the so-called α-cleavage between residues 111 and 112 in the middle of the protein (21, 22). This latter “endoproteolytic” cleavage of PrPC is reported to be carried out by members of the ADAM family (23, 24).To investigate the role of ADAMs in the ectodomain shedding of PrPC, we used loss-of-function and gain-of-function experiments in cultured cells in which candidate PrP sheddases were either knocked down by siRNA or overexpressed. We have also further characterized the shedding of BACE1 by comparison to the shedding of APP and PrPC. In addition, we have explored whether proteolytic shedding of PrPC is of importance in regulating its conversion into PrPSc.  相似文献   

13.
The biochemical essence of prion replication is the molecular multiplication of the disease-associated misfolded isoform of prion protein (PrP), termed PrPSc, in a nucleic acid-free manner. PrPSc is generated by the protein misfolding process facilitated by conformational conversion of the host-encoded cellular PrP to PrPSc. Evidence suggests that an auxiliary factor may play a role in PrPSc propagation. We and others previously discovered that plasminogen interacts with PrP, while its functional role for PrPSc propagation remained undetermined. In our recent in vitro PrP conversion study, we showed that plasminogen substantially stimulates PrPSc propagation in a concentration-dependent manner by accelerating the rate of PrPSc generation while depletion of plasminogen, destabilization of its structure and interference with the PrP-plasminogen interaction hinder PrPSc propagation. Further investigation in cell culture models confirmed an increase of PrPSc formation by plasminogen. Although molecular basis of the observed activity for plasminogen remain to be addressed, our results demonstrate that plasminogen is the first cellular protein auxiliary factor proven to stimulate PrPSc propagation.Key words: prion, PrPSc, protein misfolding, auxiliary factor, plasminogen, PMCA, cell culture modelPrions are unique infectious particles that replicate in the absence of nucleic acids1 and cause fatal neurologic disorders in mammals.2 In fact, the prion particle is composed of an alternatively folded form of the cellular prion protein (PrPC) encoded by the Prnp gene. The misfolded form of PrPC, referred to as PrPSc, shares the same primary structure with PrPC,3 but exhibits distinctively different biochemical and biophysical properties.4,5 Moreover, animals lacking expression of the Prnp gene are resistant to prion infection,6 suggesting that PrPC serves as a precursor for PrPSc.The essence of prion biosynthesis is based on the protein-only hypothesis that postulates self-perpetuating replication of the prion protein.1 Although the exact replication process remains elusive, the template-assisted conversion model proposes an idea that PrPSc serves as a template to convert α-helices of PrPC into the β-sheets of PrPSc during prion replication.7 According to many lines of evidence, there exists an unidentified auxiliary factor, designated protein X, which favorably interacts with PrPC to produce a thermodynamically stable intermediate conformation called PrP*.8 By introducing PrPSc to a host, dimers will readily form between PrP* and PrPSc. This interaction induces PrP* to take on the conformation of PrPSc and results in a complex consisting of the template and the newly formed PrPSc molecules. Once the dimer disassociates protein X and the two PrPSc molecules are released and allowed to continue replicating in an exponential fashion. Recent observations that infectious material can be generated in vitro using recombinant PrP have authenticated the protein-only hypothesis.9,10 However, the infectivity of these in vitro generated PrPSc products is lower than that of brain-derived PrPSc, leading to the possibility that insufficient levels of the unidentified auxiliary factor are the limiting factor for these in vitro assays.To date, non-mammalian chaperone proteins, sulfated glycans and certain polyanionic macromolecules, such as RNA, have shown to increase the level of PrPSc in several different in vitro assays.1114 However, defining these molecules as cellular auxiliary factors that promote the conversion of PrPC to PrPSc has been prevented for several reasons. First, overexpression of yeast heat shock protein Hsp 104 in transgenic mice does not modulate the incubation time of disease and PrPSc accumulation upon prion inoculation.15 Although mammalian Hsp 70 is upregulated in humans and animals with prion diseases,16,17 it participates in downregulation, but not upregulation, of PrPSc accumulation.18 Second, the effect of sulfated glycans on PrPSc formation has been inconsistent1113 and certain sulfated glycans inhibit PrPSc propagation in animals and cultured cells.1921 Lastly, although RNA is able to increase PrPSc propagation and induce the formation of PrPSc de novo from purified PrPC in the absence of a PrPSc seed,22 its specificity is in question. Thus, an auxiliary factor that positively assists PrPSc replication and is composed of a mammalian cellular protein remains to be identified.Our approach to identify an auxiliary factor relies on the idea that the auxiliary factor interacts with PrPC as proposed in the protein X hypothesis.7 Therefore, we considered PrP ligands as the best candidates for this unidentified factor. By screening a phage display cDNA expression library from ScN2a cells, we identified kringle domains of plasminogen that interact with recombinant PrP folded in an α-helical conformation (α-PrP).23 In vitro binding assays showed that interaction between plasminogen and α-PrP that represents the conformational state of PrPC was enhanced by the introduction of a dominant negative mutation and the presence of the basic N-terminal sequences in PrP.23 Interaction of PrPC and plasminogen was further confirmed by the ability of plasminogen and its kringle domains to readily interact with α-PrP.2429 However, despite greater interaction with α-PrP, plasminogen also interacted with PrP in β-sheet conformations.23 Prior to our study, it was shown that PrPSc was immunoprecipitated with beads linked to plasminogen, its first three kringle domains [K(1+2+3)], and more recently the repeating YRG motif found within the plasminogen kringle domains.3033 However, the ability of plasminogen to bind to PrPSc was dependent on conditions of the lipid rafts and plasminogen was actually associated with PrPC in the intact lipid rafts.34To determine the functional relevance of this interaction for PrPSc replication, we explored whether plasminogen enhances PrPSc propagation using cell culture models and the in vitro PrP conversion assay, termed protein misfolding cyclic amplification (PMCA).35 The addition of plasminogen in PMCA resulted in the generation of significantly more PrPSc in a concentration-dependent manner (Fig. 1A), suggesting that plasminogen stimulates the conversion of PrPC to PrPSc. Indeed, our kinetic studies showed that plasminogen accelerates the rate of PrPSc generation during the early stages of the in vitro PrP conversion reaction (reviewed in ref. 35). In contrast, the addition of plasminogen in PMCA lacking either PrPC or PrPSc failed to generate PrPSc (Fig. 1B and C). These results suggest that both PrPC and PrPSc are required for PrPSc replication stimulated by plasminogen and that plasminogen facilitates neither spontaneous PrP conversion nor PrPSc augmentation through aggregating pre-existing PrPSc. Incubation of plasminogen with pre-formed PrPSc followed by treatments with proteinase K failed to either increase or decrease the PrPSc level compared to controls (Fig. 1D and E), suggesting that plasminogen is not involved in stabilization of PrPSc, enhancement of PrPSc resistance to protease or promotion of PrPSc binding to the membrane for western blotting. The activity to stimulate PrP conversion was a specific property of plasminogen and was not shared with other proteins such as a known PrP ligand and proteins abundantly found in the serum or at the extracellular matrices where plasminogen is present (reviewed in ref. 35). In addition, we found that the ability of plasminogen to assist in PrPSc propagation is preserved in its kringle domains (reviewed in ref. 35). Furthermore, the activity associated with plasminogen under cell-free conditions was reproduced in cell culture models. Plasminogen and its kringle domains increased PrPSc propagation in cultured cells chronically infected with mouse-adapted scrapie or chronic wasting disease prions (Fig. 2A–D). This suggests that the activity of plasminogen in PrPSc replication has biological relevance.Open in a separate windowFigure 1The role of plasminogen in PrPSc propagation. The effect of plasminogen (Plg) was assessed by PMCA using normal brain material supplemented with or without 0.5 µM human Glu-Plg (A–D) or using Plg-deficient (Plg-/-) brain material (F and G). Pre- (−) and post- (+) PMCA samples were treated with proteinase K (PK) and analyzed by western blotting. Seeds for PMCA were diluted either as indicated or 1:900 (G) −8,100 (F). (A) Stimulation of PrPSc propagation by Plg. (B) Plg-supplemented PMCA in the absence of SBH seeds. (C) Plg-supplemented PMCA in the absence of NBH. (D) Comparison of PrPSc levels in Plg-supplemented PMCA samples (during) vs. PMCA samples only incubated with Plg prior to PK digestion (after). (E) Comparison of PrPSc levels of ScN2a cell lysate after incubation with or without Plg prior to PK digestion. (F) PMCA with brain material of Plg-/- mice and genetically unaltered littermate controls (C). (G) Restoration of PMCA using Plg-/- brain material with Plg-supplementation. NBH, normal brain homogenate; SBH, sick brain homogenate; PrPKOBH, brain homogenate of PrPC-deficient mice; CL, cell lysate. Reproduced with permission from The FASEB Journal, Mays and Ryou 2010.35Open in a separate windowFigure 2PrPSc propagation increased by plasminogen in prion-infected cells. (A) The levels of PrP in ScN2a cells incubated with 0–0.5 µM human Glu-plasminogen (Plg) for two days. (B) The levels of PrP in ScN2a cells incubated with 0, 0.1 and 1.0 µM Plg or the first three kringle domains of Plg [K(1+2+3)] for six days. (C) The levels of 3F4-tagged PrPC and nascent PrPSc formation in ScN2a cells transiently transfected (Tfx) with plasmids encoding the 3F4-tagged PrP gene (PrP-3F4) and with an empty vector (mock). The transfected cells were treated with 0 or 1 µM K (1+2+3) for three days. (D) The levels of PrP in Elk21+ cells incubated with 0 and 0.5 µM Plg for two days. PrP was detected by anti-PrP antibody D13 (A and B), 3F4 (C) or 6H4 (D) before (−) and after (+) PK treatment. Reproduced by permission of the The FASEB Journal, Mays and Ryou 2010.35Corresponding to the results that plasminogen positively assists in PrPSc replication by stimulating conversion of PrPC to PrPSc, depletion of plasminogen from the PMCA reaction by using brain material derived from plasminogen-deficient mice restricted PrPSc replication to the basal level (Fig. 1F). Supplementation with plasminogen for PMCA using plasminogen-deficient brain homogenate restored PrPSc propagation to levels equivalent to that of control PMCA in which only brain material of their non-genetically altered littermates was used (Fig. 1G). Furthermore, structural destabilization of plasminogen affected the activity of plasminogen that enhances PrPSc propagation. Because intact disulfide bonds are critical in maintaining structural integrity and the binding activity of plasminogen, we conducted PMCA supplemented with either structurally intact or modified plasminogen to investigate the functionality of plasminogen. The result showed that plasminogen pre-treated sequentially with chemical agents that disrupt disulfide bonds and modify free sulfhydryl groups failed to stimulate PrPSc propagation (reviewed in ref. 35). In addition, we showed that interference with the plasminogen-PrP interaction using L-lysine abolished the plasminogen-mediated stimulation of PrPSc propagation in PMCA (reviewed in ref. 35). Because previous observations in immunoprecipitation30 and ELISA binding assays23 described that L-lysine specifically inhibited formation of the PrP-plasminogen complex, the presence of L-lysine in PMCA is considered to saturate the lysine binding motifs of kringle domains, which competitively prevents the kringle domains of plasminogen from interacting with PrP and inhibited PrP conversion. These various inhibition studies of PrPSc propagation provides confirmatory evidence that plasminogen plays an important role in PrPSc replication.Despite our progress in understanding the role of plasminogen in PrPSc propagation, we are still unable to address mechanistic details by which plasminogen exerts its function. In fact, plasminogen shares a number of the expected characteristics of the previously proposed auxiliary factor, although there are minor but distinctive discrepancies in their properties (summarized in Fig. 3). First, plasminogen may control conformational rearrangement of PrPC to PrP*, resembling a molecular chaperone. This scenario is identical to the protein X hypothesis, in which plasminogen replaces protein X to assist the conversion process. Second, plasminogen may promote aggregation of PrPSc already converted from PrPC. This will result in efficient formation of PrPSc multimers. Third, plasminogen may stabilize the pre-existing PrPSc aggregates so that stabilized aggregates are better protected from degradation or processing by the intrinsic clearance mechanism. This will result in increased accumulation and stability of PrPSc. In the second and third scenarios, the function of plasminogen is not involved in PrPC or its conversion process, but limited to the interaction with pre-formed PrPSc. Fourth, plasminogen may play a role as a scaffolding molecule that simply brings both PrPC and PrPSc together within a proximity. This will increase the frequency of interaction between PrPC and PrPSc for conversion. This scenario is distinguished from the other three by postulating plasminogen interaction with both isoforms of PrP. In contrast, plasminogen is assumed to interact with only PrPC or PrPSc in other cases. Although accumulating data including our recent studies provide critical pieces of evidence to envision described mechanistic insights, it is still premature to conclude the mechanism involved in plasminogen-mediated stimulation of PrPSc propagation.Open in a separate windowFigure 3Plausible mechanisms for plasminogen to enhance PrPSc propagation. Plasminogen may stimulate PrPSc propagation via conformational alteration of PrPC to PrP* (i), enhancement of PrPSc aggregation (ii), stabilization of pre-exisiting PrPSc aggregates (iii) or scaffolding to gather PrPC and PrPSc together (iv).

Table 1

Properties of plasminogen as an auxiliary factor for PrPSc propagation
Protein XPlasminogen
CompositionProtein; macromoleculesProtein
ExpressionBrain; neuron-specificBrain; neuroblastoma cell line, expressed more in the non-CNS
Subcellular localizationPlasma membrane; lipid raftsExtracellular matrix; lipid rafts
Association with diseaseIncreased protein levels in the sera of human patient with CJD
InteractionOnly with PrPCPrPSc, α-PrP, β-PrP
Binding sites on PrP
  • Q167, Q171, V214 and Q218 in the β2-α2 loop (164–174) and C-terminus (215–223)
  • Dominant negative mutations on the protein X binding sites such as Q167R and Q218K inhibited PrPSc formation in the cultured cells and prion transmission in transgenic mice
  • K23, K24 and K27 in the N-terminus: deletion of the N-terminal lysine cluster reduced dominant negative inhibition of PrPSc formation
  • Binds to N-terminally truncated PrP (89–230)
  • Increased binding activity to full-length PrP (23–230)
  • Increased binding activity to PrP with Q218K dominant negative mutation
  • The second kringle domain of plasminogen binds to the β2-α2 loop in silico
  • Binds to both lysine clusters located to the N-terminus (23–27) and middle (100–109) of PrP
Species specificity
  • Homotypic interaction with PrPC
  • Mouse protein X has lower binding affinity to human PrPC in the studies with transgenic mice
  • Unknown
  • Human plasminogen binds to human, mouse, bovine and ovine PrP
  • Human and bovine plasminogen converts mouse PrPC to PrPSc in PMCA
FunctionAn auxiliary role in conversion of PrPC to PrPScEnhances PrPSc propagation facilitated by PrP conversion in PMCA
Action mechanismBinds to PrPC and alters PrPC into PrP* that interacts with PrPSc for conversionUnknown
Open in a separate windowCNS, central nervous system; CJD, Creutzfeldt-Jakob disease; α-PrP, PrP in an α-helical conformation; β-PrP, PrP in an β-sheet conformation; PMCA, protein misfolding cyclic amplification.It is essential to address a few additional issues of PrPSc propagation stimulated by plasminogen aside from the mechanistic details. First, it is necessary to confirm the authenticity of PrPSc generated in the aid of plasminogen as a bona fide infectious agent. In addition, it would be interesting to compare the structural signatures of PrPSc generated in our study and found the prion seeds. This study may provide a clue to establish a structure-infectivity relationship for PrPSc. Furthermore, the ability to repeat our PMCA results in a similar assay reconstituted with defined components would clarify whether plasminogen directly contributes to PrPSc propagation. Finally, the contribution of plasminogen to the species barrier by controlling the compatibility between the host PrPC and prion strains should be determined. Collectively, these studies will be useful to identify an intricate regulatory role for plasminogen during PrPSc replication and prion transmission.Plasminogen has shown that it stimulates PrPSc propagation in the cell-free assays and cultured cells, while its role in animal models has not been obviously clarified. In opposition to the anticipated outcomes that the course of prion disease in plasminogen-deficient animals would be delayed, intracerebral prion infection of two independent mouse lines deficient in plasminogen resulted in either unchanged or accelerated disease progression when compared to the appropriate wild-type controls.36,37 These results do not support each other and there is no good interpretation to reconcile their incongruence. However, the discrepancy between the expectation and the actual outcomes of the previous studies can be caused by the intrinsic health problems associated with the mouse lines used for the studies,3840 which raises a question for the credibility of the model system. Some clinical phenotypes of confounding health problems in plasminogen-deficient mice overlap with the typical clinical signs of prion disease in mice. Furthermore, the drastically shortened life expectancy of plasminogen-deficient mice coincides with the incubation periods of mice inoculated with prion strains used in the previous studies. Lastly, the health problems of plasminogen-deficient mice could exacerbate the progression of the course of disease in mice inoculated with prions. Thus, either unchanged or shortened incubation periods of prion-inoculated plasminogen-deficient mice may not solely reflect the effect of plasminogen deficiency related to prion replication. Alternatively, the reason for observed outcomes from the in vivo models may be associated with the presence of functionally redundant proteins for plasminogen even under its absence. Because kringle domains of plasminogen share a well-conserved structure with those of other proteins such as tissue-type plasminogen activator, hepatocyte growth factor and apolipoprotein (a),41 it is possible to postulate that other proteins that contain kringle domains can functionally replace plasminogen. Therefore, an animal model that circumvents the drawbacks associated with the current models is needed to address the relevance of plasminogen in prion replication in vivo.Based on our study that identified plasminogen as the first cellular protein cofactor for PrPSc propagation, we anticipate an intriguing opportunity to develop future diagnosis and therapeutic intervention for prion disease. Previously, the identity of a proposed auxiliary factor was obscure so most approaches either targeted PrP isoforms or were empirical.42 Our study suggests that plasminogen is a novel target to interfere with PrPSc replication. Therefore, a variety of strategies that either deplete plasminogen or interfere with the formation of PrP-plasminogen complexes would work for development of a novel therapeutic intervention for prion disease. For instance, RNA interference of plasminogen expression, monoclonal anti-plasminogen antibody to block plasminogen binding to PrP, L-lysine that saturates the binding sites of plasminogen to PrP, and chemical agents that destabilize the structure of plasminogen are potential strategies for this purpose.In conclusion, plasminogen is a PrP ligand with the ability to stimulate PrPSc propagation. Our findings are indispensable in gaining a better understanding of the underlying mechanism for PrPSc propagation, while unveiling a new therapeutic target for prion disease.  相似文献   

14.
15.
The conversion to a disease-associated conformer (PrPSc) of the cellular prion protein (PrPC) is the central event in prion diseases. Wild-type PrPC converts to PrPSc in the sporadic forms of the disorders through an unknown mechanism. These forms account for up to 85% of all human (Hu) occurrences; the infectious types contribute for less than 1%, while genetic incidence of the disease is about 15%. Familial Hu prion diseases are associated with about 40 point mutations of the gene coding for the PrP denominated PRNP. Most of the variants associated with these mutations are located in the globular domain of the protein. In a recent work in collaboration with the German Research School for Simulation Science, in Jülich, Germany, we performed molecular dynamics simulations for each of these mutants to investigate their structure in aqueous solution. Structural analysis of the various point mutations present in the globular domain unveiled common folding traits that may allow a better understanding of the early conformational changes leading to the formation of monomeric PrPSc. Recent experimental data support these findings, thus opening novel approaches to determine initial structural determinants of prion formation.Key words: prions, prion protein, human, pathogenic mutations, structure, molecular dynamics, nuclear magnetic resonancePrion diseases have attracted much attention from researchers with different scientific backgrounds and coming from various areas of expertise. Many questions still remain unanswered in the study of these rare and yet unique neurodegenerative disorders. Central to understanding the disease is deciphering the nature of the causative agent of these disorders: the prion. In fact, the mechanism by which a prion (PrPSc) is formed and the structure of the latter, have posed major challenges to this field. Indeed, prion research has achieved a great deal of detailed information in understanding the pathogenesis of the disease, but until now the early events leading to the conformational change harboring prions have remained elusive.1 In an attempt to learning how the protein may undergo this conformational rearrangement, my group and the group of Paolo Carloni at the German Research School for Simulation Science, in Jülich, Germany, reasoned that some clues might come from the study of pathogenic mutants in HuPrP. At the time of beginning our work the structures of few mutants were known.2 The structure of HuPrP was used as template for our studies.3 We therefore performed molecular dynamics (MD) simulations for each of these mutants to investigate their structure in aqueous solution. In total, almost 2 µs MD data were obtained. The calculations were based on the AMBER(parm99) force field, which has been shown to reproduce very accurately the structural features of the wild-type HuPrP and a few variants for which experimental structural information was available.4 All the variants present structural features different from those of wild-type HuPrP and the protective dominant negative polymorphism HuPrP(E219K). These characteristics include loss of salt bridges in the α2–α3 regions and the loss of π-stacking interactions in the β22 loop. In addition, in the majority of the mutants analyzed, the α3 helix is more flexible and the residue Tyr169 is more exposed to the aqueous solvent. The biological relevance of these findings is of utmost importance. The presence of similar traits in this large spectrum of mutations hints to a role of these characteristics in their known capabilities to generate disease-causing properties. Overall, we concluded that the regions most affected by disease-linked mutations in terms of structure and/or flexibility might be those involved in the pathogenic conversion of PrPC to the scrapie form of the protein, and ultimately, in the interaction with cellular partners.Recent reports have indicated that the alteration of PRNP sequence by pathological mutations is sufficient to generate prions in transgenic mice.5 Therefore, solution-state NMR studies on PrP mutants may help identifying critical regions in PrPC structure involved in the conversion. The comparison between the structures of Q212P and V210I mutants with the wild-type HuPrP revealed that, although structures share similar globular architecture, mutations introduce novel local structural differences.6 The observed variations are mostly clustered in the β22-loop region and in the α2–α3 inter-helical interfaces. In contrast to the wild-type protein, where the structures of Q212P and V210I mutants point to the interruption of aromatic and hydrophobic interactions between residues located at the interface of the β22 loop and the C-terminal end of α3 helix. A loss of contacts between the β22-loop and the α3 helix in the mutants results in higher exposure of hydrophobic residues to solvent. Similar findings have also been reported in the NMR structure of the E200K mutant,7 X-ray structures of F198S and D178N mutants8 and in independent MD studies.911 In addition, in the two mutants here considered side chains of Phe141 and Tyr149 adopt different orientation. Our findings indicate that the structural disorder of the β22-loop together with the increased distance between the loop and α3 helix represent key pathological structural features and may shed light on critical epitopes on the HuPrP structure possibly involved in the conversion to PrPSc.Different experimental studies suggested that the conformation of the β22-loop plays a role in the prion disease transmission and susceptibility. Several studies have indicated that mammals carrying a flexible β22 loop could be easily infected by prions, whereas prions are poorly transmissible to animals carrying a rigid loop.12 Importantly, horse and rabbit have so far displayed resistance to prion infections and there are no reports of these species developing spontaneous prion diseases. NMR studies showed that their PrP structures are characterized by a rigid β22 loop and by closer contacts between the loop and α3 helix.13,14 Thus, it seems that prion resistance is enciphered by the amino acidic composition of the β22-loop and its long-range interactions with the C-terminal end of the α3 helix.Interestingly, it has been proposed a role of α1 helix as a promoter of PrPC aggregation.15 In support, Tyr149 in α1 helix is part of a motif, which may be solvent exposed in PrPSc and involved in structural rearrangements during fibril formation.16 Pronounced stabilization of α1 helix in the protein may represent another important factor in the prevention of spontaneous PrPSc formation.Comparing the structures of the wild-type protein and the mutants enabled us to detect regions on HuPrP structure that may play a key role in the pathogenic conversion. The obtained structural data indicate that the β22 loop and, in particular, interactions of this loop with residues in the C-terminal part of α3 helix determine the extent of exposure of hydrophobic surface to solvent, and thus could influence propensity of PrPC for intermolecular interactions. Moreover, our results highlight the significance of the α1 helix and its tertiary contacts in overall stabilization of HuPrP folding.Overall, the many features discussed here involve the most important regions that confer stability to wild-type HuPrP, although the mutations considered are different for position and characteristics. In particular, the β22-loop and the α2–α3 regions are the most affected in terms of structural organization and flexibility of the molecule. These two subdomains are crucial for the stability of the wild-type HuPrPC fold17 and might play a prominent role in the early unfolding events leading to PrPSc conversion.  相似文献   

16.
It is now well established that the conversion of the cellular prion protein, PrPC, into its anomalous conformer, PrPSc, is central to the onset of prion disease. However, both the mechanism of prion-related neurodegeneration and the physiologic role of PrPC are still unknown. The use of animal and cell models has suggested a number of putative functions for the protein, including cell signaling, adhesion, proliferation, and differentiation. Given that skeletal muscles express significant amounts of PrPC and have been related to PrPC pathophysiology, in the present study, we used skeletal muscles to analyze whether the protein plays a role in adult morphogenesis. We employed an in vivo paradigm that allowed us to compare the regeneration of acutely damaged hind-limb tibialis anterior muscles of mice expressing, or not expressing, PrPC. Using morphometric and biochemical parameters, we provide compelling evidence that the absence of PrPC significantly slows the regeneration process compared to wild-type muscles by attenuating the stress-activated p38 pathway, and the consequent exit from the cell cycle, of myogenic precursor cells. Demonstrating the specificity of this finding, restoring PrPC expression completely rescued the muscle phenotype evidenced in the absence of PrPC.The cellular prion protein (PrPC) is a glycoprotein, prominently expressed in the mammalian central nervous system (CNS) and lymphoreticular system, that is anchored to the cell external surface through a glycolipidic moiety. The bad reputation acquired by PrPC originates from the notion that an aberrant conformer of it (PrPSc) is the major component of the prion, the unconventional infectious particle that causes fatal neurodegenerative disorders, i.e., transmissible spongiform encephalopathies (TSE) or prion diseases (56). A wealth of evidence has suggested that the function of PrPC is beneficial to the cell, but currently, our detailed comprehension of its physiology remains poor. In this respect, the availability of knockout (KO) paradigms for PrPC has provided less crucial information than expected. Subtle phenotypes, e.g., mild neuropathologic, cognitive, and behavioral deficits, have been described in PrP-KO mice (17, 50), but these animals generally live a normal life span without displaying obvious developmental defects (8, 42). Importantly, the same holds true when the expression of PrPC is postnatally abrogated (40). The extensive search for PrPC''s raison d''être has ascribed to the protein a plethora of functions (for updated reviews, see references 1 and 35); among these, roles in cell adhesion, migration, and differentiation have been proposed whereby PrPC could act by modulating different cell-signaling pathways (63). In this framework, a variety of neuronal proteins have been hypothesized to interact with PrPC (reviewed in references 1 and 11), for example, cell adhesion molecules or extracellular matrix proteins, which could explain the capacity of PrPC to mediate the neuritogenesis and neuronal differentiation observed in several cell model systems (13, 22, 23, 27, 36, 59, 64).Although neurons are generally regarded as the model of choice for unraveling the function of PrPC, the expression of the protein in several other organs suggests that PrPC has a conserved role in different tissues. Thus, important insight into PrPC function may also be provided by the analysis of extraneural tissues. One such tissue is skeletal muscle, which has been shown to express PrPC at significant levels (43, 46) and has been found to upregulate PrPC levels under stress conditions (71). On the other hand, ablation of the PrP gene has been shown to directly affect skeletal muscles, for example, by enhancing oxidative damage (30) or by diminishing tolerance for physical exercise (51). Skeletal muscles have also been associated with prion pathology, as evidenced by the accumulation of PrPSc (or PrPSc-like forms) in the muscles of TSE-affected humans and animals (2, 3, 6, 21, 53, 67) and by transgenic-mouse models of some inherited TSEs (16). In addition, overexpression of wild-type (WT) PrPC (25, 68), or expression of TSE-associated mutants of the protein (16, 66), generates myopathic traits in transgenic mice.In light of these notions, and because intact muscle tissues are more amenable to in vivo manipulations than neural tissue, we set out to analyze the potential role of PrPC in tissue morphogenesis (38, 41, 46) using an in vivo skeletal-muscle paradigm from two congenic mouse lines expressing (WT) or not expressing (PrP-KO) PrPC. Importantly, to verify that the PrP-KO muscle phenotype was specifically dependent on the absence of PrPC, we used PrP-KO mice reconstituted with a PrP transgene (PrP-Tg). The applied protocol consisted of first characterizing the degeneration of the hind-limb tibialis anterior (TA) muscle and then evaluating the myogenic process from the response to inflammation to the full recovery of the muscle. By combining acute insult with adult age, this strategy also had the potential to bypass possible compensatory mechanisms that might mask PrP-KO phenotypes during embryogenesis and/or in adulthood under normal conditions (65).In this study, we provide evidence that, compared to animals expressing PrPC (WT and PrP-Tg), recovery from damage of adult skeletal muscles was significantly slower in PrP-KO mice. Analysis of the different stages of muscle regeneration allowed us to conclude that PrPC is one of the factors that govern the early phases of this process, in which the proliferation and differentiation of myogenic precursor cells take place.  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.
Although cellular prion protein (PrPc) has been suggested to have physiological roles in neurogenesis and angiogenesis, the pathophysiological relevance of both processes remain unknown. To elucidate the role of PrPc in post-ischemic brain remodeling, we herein exposed PrPc wild type (WT), PrPc knockout (PrP−/−) and PrPc overexpressing (PrP+/+) mice to focal cerebral ischemia followed by up to 28 days reperfusion. Improved neurological recovery and sustained neuroprotection lasting over the observation period of 4 weeks were observed in ischemic PrP+/+ mice compared with WT mice. This observation was associated with increased neurogenesis and angiogenesis, whereas increased neurological deficits and brain injury were noted in ischemic PrP−/− mice. Proteasome activity and oxidative stress were increased in ischemic brain tissue of PrP−/− mice. Pharmacological proteasome inhibition reversed the exacerbation of brain injury induced by PrP−/−, indicating that proteasome inhibition mediates the neuroprotective effects of PrPc. Notably, reduced proteasome activity and oxidative stress in ischemic brain tissue of PrP+/+ mice were associated with an increased abundance of hypoxia-inducible factor 1α and PACAP-38, which are known stimulants of neural progenitor cell (NPC) migration and trafficking. To elucidate effects of PrPc on intracerebral NPC homing, we intravenously infused GFP+ NPCs in ischemic WT, PrP−/− and PrP+/+ mice, showing that brain accumulation of GFP+ NPCs was greatly reduced in PrP−/− mice, but increased in PrP+/+ animals. Our results suggest that PrPc induces post-ischemic long-term neuroprotection, neurogenesis and angiogenesis in the ischemic brain by inhibiting proteasome activity.Endogenous neurogenesis persists in the adult rodent brain within distinct niches such as the subventricular zone (SVZ) of the lateral ventricles,1, 2, 3, 4 which host astrocyte-like neural stem cells and neural progenitor cells (NPCs). Focal cerebral ischemia stimulates neurogenesis, and NPCs proliferate and migrate towards the site of lesion where they eventually differentiate.5, 6, 7 In light of low differentiation rates and high cell death rates of new-born cells,6, 8, 9 post-stroke neurogenesis is scarce.10Cellular prion protein (PrPc) is a glycoprotein that is attached to cell membranes by means of a glycosylphosphatidylinositol anchor.11 Although PrPc is ubiquitously expressed, it is most abundant within the central nervous system. Conversion into its misfolded isoform PrPsc causes neurodegenerative diseases such as Creutzfeldt-Jacob disease.11, 12 While a large body of studies analyzed the role of PrPsc in the context of transmissible spongiform encephalopathies, little is known about the physiological role of PrPc. Studies performed during both ontogenesis and adulthood suggest that PrPc regulates neuronal proliferation and differentiation, synaptic plasticity and angiogenesis.13, 14, 15, 16, 17, 18 The role of these processes under pathophysiological conditions, however, is largely unknown.Previous reports suggested a role of PrPc in post-ischemic neuroprotection.19, 20, 21, 22, 23, 24 Thus, PrPc was found to be overexpressed in ischemic brain tissue.19, 20, 21, 22, 23, 24 PrPc deficiency aggravated ischemic brain injury, possibly via enhanced ERK-1/2 activation and reduced phosphorylation of Akt, thus ultimately culminating in increased caspase-3 activity,21, 24 whereas PrPc overexpression protected against ischemia.19, 20, 21, 22, 23, 24 Nevertheless, these studies focused on acute injury processes with a maximal observation period of 3 days, leaving the biological role of PrPc in post-stroke neurogenesis and angiogenesis unanswered. To clarify the role of PrPc in the post-acute ischemic brain, we herein exposed PrPc wild type (WT), PrPc knockout (PrP−/−) and PrPc overexpressing (PrP+/+) mice to focal cerebral ischemia induced by intraluminal middle cerebral artery (MCA) occlusion, evaluating effects of PrPc on neurological recovery, ischemic injury, neurogenesis and angiogenesis, as well as the homing and efficacy of exogenously delivered NPCs.  相似文献   

20.
Normal cellular and abnormal disease-associated forms of prion protein (PrP) contain a C-terminal glycophosphatidyl-inositol (GPI) membrane anchor. The importance of the GPI membrane anchor in prion diseases is unclear but there are data to suggest that it both is and is not required for abnormal prion protein formation and prion infection. Utilizing an in vitro model of prion infection we have recently demonstrated that, while the GPI anchor is not essential for the formation of abnormal prion protein in a cell, it is necessary for the establishment of persistent prion infection. In combination with previously published data, our results suggest that GPI anchored PrP is important in the amplification and spread of prion infectivity from cell to cell.Key words: prion, GPI anchor, PrP, prion spread, scrapieIn transmissible spongiform encephalopathies (TSE or prion diseases) such as sheep scrapie, bovine spongiform encephalopathy and human Creutzfeldt-Jakob disease, normally soluble and protease-sensitive prion protein (PrP-sen or PrPC) is converted to an abnormal, insoluble and protease-resistant form termed PrP-res or PrPSc. PrP-res/PrPSc is believed to be the main component of the prion, the infectious agent of the TSE/prion diseases. Its precursor, PrP-sen, is anchored to the cell surface at the C-terminus by a co-translationally added glycophosphatidyl-inositol (GPI) membrane anchor which can be cleaved by the enzyme phosphatidyl-inositol specific phospholipase (PIPLC). The GPI anchor is also present in PrP-res, but is inaccessible to PIPLC digestion suggesting that conformational changes in PrP associated with PrP-res formation have blocked the PIPLC cleavage site.1 Although the GPI anchor is present in both PrP-sen and PrP-res, its precise role in TSE diseases remains unclear primarily because there are data to suggest that it both is and is not necessary for PrP-res formation and prion infection.In tissue culture cells infected with mouse scrapie, PrP-res formation occurs at the cell surface and/or along the endocytic pathway24 and may be dependent upon the membrane environment of PrP-sen. For example, localization via the GPI anchor to caveolae-like domains favors PrP-res formation5 while substitution of the GPI anchor addition site with carboxy termini favoring transmembrane anchored PrP-sen inhibits formation of PrP-res.5,6 Other studies have shown that localization of both PrP-sen and PrP-res to lipid rafts, cholesterol and sphingolipid rich membrane microdomains where GPI anchored proteins can be located, is important in PrP-res formation.69However, there are also data which suggest that such localization is not necessarily essential for PrP-res formation. Anchorless PrP-sen isolated from cells by immunoprecipitation or wild-type PrP-sen purified by immunoaffinity column followed by cation exchange chromatography are efficiently converted into PrP-res in cell-free systems.10,11 Furthermore, recombinant PrP-sen derived from E. coli, which has no membrane anchor or glycosylation, can be induced to form protease-resistant PrP in vitro when reacted with prion-infected brain homogenates.1214 Finally, in at least one instance, protease-resistant recombinant PrP-res generated in the absence of infected brain homogenate was reported to cause disease when inoculated into transgenic mice.15The data concerning the role of the PrP-sen GPI anchor in susceptibility to TSE infection are similarly contradictory. Transgenic mice expressing anchorless mouse PrP-sen are susceptible to infection with mouse scrapie and accumulate both PrP-res and prion infectivity.16 Thus, the GPI anchor is clearly not needed for PrP-res formation or productive TSE infection in vivo. However, we recently published data demonstrating that, in vitro, anchored PrP-sen is in fact required to persistently infect cells.17 Given that anchorless PrP-sen is not present on the cell surface but is released into the cell medium, we speculated that the differences between the in vitro and in vivo data were related to the location of PrP-res formation. In the mice expressing anchorless PrP-sen, environments conducive to PrP-res formation are present in certain areas of the complex extracellular milieu of the brain where anchorless, secreted PrP-sen can accumulate and come into contact with PrP-res from the infectious inoculum. Since similar environments are missing in vitro, any PrP-res formation in cells expressing anchorless PrP-sen must be cell-associated. While this explanation addresses how extracellular PrP-res could be generated in an unusual transgenic mouse model of TSE infection, it does not really help to define how the GPI anchor is involved in normal prion infection of a cell.As with other infectious organisms such as viruses, TSE infection can be roughly divided into three steps: uptake, replication and spread. Over the last several years, data derived from new techniques as well as new cell lines susceptible to prion infection have increased our knowledge of some of the basic events that occur during each of these steps. In order to try to tease out the role of the GPI anchor in normal TSE pathogenesis, it is therefore useful to consider the process of TSE infection of a cell and how the GPI anchor might be involved in each stage.In a conventional viral infection, binding and uptake of the virus is essential to establish infection. Studying PrP-res uptake has been complicated by the lack of an antibody that can specifically distinguish PrP-res from PrP-sen in live cells and by the difficulty of detecting the input PrP-res from the PrP-res made de novo by the cell. Recently, however, several groups have been able to study PrP-res uptake using input PrP-res that was either fluorescently labeled1820 or tagged with the epitope to the monoclonal antibody 3F4,21 or cell lines that express little or no PrP-sen.19,2123 The data show that PrP-res uptake is independent of scrapie strain or cell type but is influenced by the PrP-res microenvironment as well as PrP-res aggregate size.21 Importantly, these studies demonstrated that PrP-sen expression was not required.19,2123 Given these data, it is clear that GPI anchored PrP-sen is not involved in the initial uptake of PrP-res into the cell.The next stage of prion infection involves replication of infectivity which is typically assayed by following cellular PrP-res formation. Once again, however, the issue of how to distinguish PrP-res in the inoculum from newly formed PrP-res in the cells has made it difficult to study the early stages of prion replication. To overcome this difficulty, we developed a murine tissue culture system that utilizes cells expressing mouse PrP-sen tagged with the epitope to the 3F4 antibody (Mo3F4 PrP-sen).24 Wild-type mouse PrP does not have this epitope. As a result, following exposure to an infected mouse brain homogenate, de novo PrP-res formation can be followed by assaying for 3F4 positive PrP-res. Our studies showed that there were two stages of PrP-res formation: (1) an initial acute burst within the first 96 hours post-infection that was cell-type and scrapie strain independent and, (2) persistent PrP-res formation (i.e., formation of PrP-res over multiple cell passages) that was dependent on cell-type and scrapie strain and associated with long-term infection.24 Acute PrP-res formation did not necessarily lead to persistent PrP-res formation suggesting that other cell-specific factors or processes are needed for PrP-res formation to persist.24When cells expressing Mo3F4 PrP-sen without the GPI anchor (Mo3F4 GPI-PrP-sen) were exposed to mouse scrapie infected brain homogenates, GPI negative, 3F4 positive PrP-res (Mo3F4 GPI-PrP-res) was detected within 96 hours indicating that acute PrP-res formation had occurred.17 Thus, despite the fact that Mo3F4 GPI-PrP-sen is not expressed on the cell surface16 (Fig. 1A), it was still available for conversion to PrP-res. These results are consistent with data from cell-free systems and demonstrate that, at least acutely, membrane anchored PrP is not necessary for PrP-res formation in a cell.Open in a separate windowFigure 1Persistent infection of cells in vitro requires the expression of GPI-anchored cell surface PrP-sen. PrP knockout cells (CF10)21 were transduced with 3F4 epitope tagged mouse PrP-sen (Mo3F4), 3F4 epitope tagged mouse PrP-sen without the GPI anchor (Mo3F4 GPI-), or Mo3F4 GPI-PrP-sen plus wild-type, GPI anchored mouse PrP-sen (MoPrP). The cells were then exposed to the mouse scrapie strain 22L and passaged. (A) The presence of 3F4 epitope tagged, cell surface mouse PrP-sen was assayed by FACS analysis of fixed, non-permeabilized cells. CF10 cells expressing the following mouse PrP-sen molecules were assayed: Mo3F4 (solid line); Mo3F4 GPI (dashed line); Mo3F4 GPI + MoPrP (dotted and dashed line); Mo3F4 GPI + MoPrP infected with 22L scrapie (dotted line). Only cells expressing Mo3F4 PrP-sen were positive for cell surface, 3F4 epitope tagged PrP. (B) Persistent infection was analyzed by inoculating the cells intracranially into transgenic mice overexpressing MoPrP (Tga20 mice). Only cells expressing anchored mouse PrP-sen were susceptible to scrapie infection. Cells expressing anchorless mouse PrP-sen did not contain detectable infectivity in either the cells or the cellular supernatant (data not shown). Data in (B) are adapted from McNally 2009.17In terms of persistent PrP-res formation, however, our data suggest that the GPI anchor is important. Despite an initial burst of PrP-res formation within the first 96 hours post-infection, Mo3F4 GPI-PrP-res was not observed following passage of the cells nor did the cells become infected. This effect was not due either to resistance of the cells to scrapie infection or to an inability of the scrapie strain used to infect cells. When the same cells expressed anchored Mo3F4 PrP-sen and were exposed to the same mouse scrapie strain, both acute and persistent PrP-res formation were detected and the cells were persistently infected with scrapie (Fig. 1B).17 Taken together, these data demonstrate that cells expressing anchorless PrP-sen do not support persistent PrP-res formation. Furthermore, the data strongly suggest that GPI-anchored PrP-sen is required during the transition from acute to persistent scrapie infection. In support of this hypothesis, the resistance of cells expressing Mo3F4 GPI-PrP-sen to persistent prion infection could be overcome if wild-type GPI anchored PrP-sen was co-expressed in the same cell. When both forms of PrP-sen were expressed, anchored and anchorless forms of PrP-res were made and the cells became persistently infected (Fig. 1B).17 Thus, the data suggest that GPI anchored PrP is necessary to establish prion infection within a cell.How could GPI membrane anchored PrP be involved in the establishment and maintenance of persistent prion infection? Several studies have suggested that the GPI anchor is needed to localize PrP-sen to specific membrane environments where PrP-res formation is favored.58 However, if this localization was essential for PrP-res formation, GPI-PrP-sen would presumably never form PrP-res. Lacking the GPI anchor, it would not be in the correct membrane environment to support conversion. As a result, neither acute nor persistent prion infection could occur. This is obviously not the case. Transgenic mice expressing only anchorless PrP-sen generate PrP-res and can be infected with scrapie even though (1) flotation gradients showed that anchorless PrP-sen was not in the same membrane environment as anchored PrP-sen and, (2) flow cytometry analysis demonstrated that anchorless PrP-sen was not present on the cell surface.16 Thus, the GPI anchor is not needed to target PrP-sen to a conversion friendly membrane environment.Consistent with the idea that the GPI anchor is not essential for PrP-res formation, in our studies anchorless PrP-sen could form PrP-res in cells acutely infected with scrapie despite the fact that it is processed differently than anchored PrP-sen, is not present on the cell surface (Fig. 1A), and is secreted.17 Persistent formation of anchorless PrP-res only occurred when both anchored and anchorless forms of PrP were expressed in the same cell.17 For this to happen both types of PrP must share a cellular compartment where PrP-res formation occurs, presumably either on the cell surface or in a specific location along the endocytic pathway2,3 such as the endosomal recycling compartment.4 Analysis of infected and uninfected cells co-expressing Mo3F4 GPI-PrP-sen and wild-type PrP-sen demonstrated that Mo3F4 GPI-PrP-sen was not present on the cell surface (Fig. 1A). Thus, it is unlikely that GPI-PrP-res formation is occurring on the cell surface. We speculate that the anchored form of PrP-res encounters anchorless PrP-sen along either a secretory or endocytic pathway, allowing for the formation of anchorless PrP-res. Regardless of the precise location, the in vitro and in vivo data strongly suggest that the role of the anchor in persistent prion infection is not simply to localize PrP-sen to an environment compatible with PrP-res formation.However, the data are consistent with the idea that GPI anchored PrP is absolutely essential for the establishment of persistent infection in vitro. This is likely related to the spread of infectivity within a culture that is necessary for maintaining a persistent infection over time. Evidence suggests that PrP-res can be transferred between cells in a variety of ways including mother-daughter cell division,25 cell-to-cell contact26,27 and exosomes.28 Tunneling nanotubes have also been hypothesized to be involved in intercellular prion spread19 and recent data suggest that spread can occur via these structures.20 Any of these processes could involve the cell-to-cell transfer of PrP-res in membrane containing particles as has been observed in cell-free7 and cell-based systems.29 If cell-to-cell contact were required, for example via simple physical proximity or perhaps tunneling nanotubes,19,20 then the conversion of cell surface PrP-sen on the naïve cell by cell surface PrP-res on the infected cell would transfer infection to the naïve cell. In this instance, GPI membrane anchored, cell surface PrP-sen would be essential as it would allow for PrP-res formation on the cell surface. If spread is via cell division, then GPI-anchored, cell surface PrP-sen would be important for its role as a precursor to PrP-res formation.2 In this instance, cell surface PrP-sen would be an essential intermediate in the continuous formation of PrP-res necessary for the accumulation and amplification of PrP-res within the cell. It would also help to cycle PrP between the cell surface and intracellular compartments where PrP-res can be formed.4 In either case, GPI-anchored PrP-sen would facilitate the accumulation of intracellular PrP-res to high enough levels to maintain both persistent infection in the mother cell and enable the transfer of organelles containing sufficient PrP-res to initiate infection in the daughter cell. Thus, we would suggest that efficient spread of infectivity requires not just the passive transfer of PrP-res from cell-to-cell but the concurrent initiation of conversion and amplification of PrP-res via cell surface, GPI anchored PrP-sen.In vivo, several lines of evidence suggest that the spread of scrapie infectivity also requires de novo PrP-res formation in the recipient cell and not simply transfer of PrP-res from one cell to another. For example, when neurografts from PrP expressing mice were placed in the brains of PrP knockout mice and the mice were challenged intracranially with scrapie, the graft showed scrapie pathology, but the surrounding tissue did not.30 Furthermore, PrP-res from the graft migrated to the host tissue demonstrating that simple transfer of PrP-res was not sufficient and that PrP-sen expression was required for the spread of scrapie pathology.30 In fact, these mice did not develop scrapie pathology following peripheral infection even when peripheral lymphoid tissues were reconstituted with PrP-sen expressing cells.31 Even though PrP-sen expressing cells were present in both the brain and spleen, in order for infectivity to spread from the lymphoreticular system to the central nervous system PrP-sen expression was also required in an intermediate tissue such as peripheral nerve.31,32 Given that PrP-res uptake and trafficking do not require PrP-sen, the most obvious explanation for the requirement of PrP-sen in contiguous tissues is that de novo PrP-res formation in naïve cells is necessary for (1) infectivity to move from cell to cell within a tissue and, (2) infectivity to move from tissue to tissue.Another study demonstrated that peripheral expression of heterologous mouse PrP significantly increased the incubation time and actually prevented clinical disease in the majority of transgenic mice expressing hamster PrP in neurons of the brain.33 Once again, if simple transfer and uptake of PrP-res were sufficient for spread, the presence of heterologous PrP molecules should not interfere because cellular uptake of PrP-res is independent of PrP-sen expression.19,2123 Clinical disease in these mice was likely prevented by the heterologous PrP molecule interfering with conversion of PrP-sen to PrP-res suggesting that prevention of de novo PrP-res formation inhibits spread of PrP-res and infectivity. These in vivo data, when combined with our recent in vitro data,17 provide evidence to support the importance of cell surface, and by extension GPI-anchored, PrP in the spread of prion infection.Our data demonstrate that the GPI anchor plays a role in the establishment of persistent scrapie infection in vitro. In our tissue culture system,21 as well as others where spread of infectivity by cell to cell contact appears to be limited,25,34 the role of GPI anchored PrP-sen would be to amplify PrP-res to enable the efficient transfer of infectivity from mother to daughter cell. In cell systems where spread of prion infectivity may require cell to cell contact,26,27 we propose that the role of GPI anchored PrP-sen is to facilitate the spread of prion infection via a chain of conversion from cell-to-cell, a “domino” type spread of infection that has been previously hypothesized.35,36In vivo, such a mechanism might explain why neuroinvasion does not necessarily require axonal transport32,37,38 and can occur independently of the axonal neurofilament machinery.39 It would likely vary with cell type27 and be most important in areas where infectivity is transferred from the periphery to the nervous system as well as in areas where cell division may be limited. It is also possible, if the location of PrP-res formation differs for different scrapie strains,40 that the relative importance of a domino-like spread of infectivity in vivo would vary with the scrapie strain.Of course, spread of infectivity via a “wave” of GPI anchored, PrP mediated conversion would not preclude the spread of infectivity by other intracellular means such as axonal transport (reviewed in ref. 41). Furthermore, spread of infectivity may still also occur extracellularly such as in the unique case of mice which express anchorless PrP-sen,16 where our in vitro data would suggest that the cells themselves are not infected. In such a case, spread would require neither GPI anchored PrP-sen nor amplification of PrP-res in cells but would likely occur via other means such as blood41 or interstitial fluid flow.42  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