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 46 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.
4.
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   

5.
6.
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   

7.
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   

8.
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   

9.
The clinicopathological phenotypes of sporadic Creutzfeldt-Jakob disease (sCJD) correlate with the allelotypes (M or V) of the polymorphic codon 129 of the human prion protein (PrP) gene and the electrophoretic mobility patterns of abnormal prion protein (PrPSc). Transmission of sCJD prions to mice expressing human PrP with a heterologous genotype (referred to as cross-sequence transmission) results in prolonged incubation periods. We previously reported that cross-sequence transmission can generate a new prion strain with unique transmissibility, designated a traceback phenomenon. To verify experimentally the traceback of sCJD-VV2 prions, we inoculated sCJD-VV2 prions into mice expressing human PrP with the 129M/M genotype. These 129M/M mice showed altered neuropathology and a novel PrPSc type after a long incubation period. We then passaged the brain homogenate from the 129M/M mouse inoculated with sCJD-VV2 prions into other 129M/M or 129V/V mice. Despite cross-sequence transmission, 129V/V mice were highly susceptible to these prions compared to the 129M/M mice. The neuropathology and PrPSc type of the 129V/V mice inoculated with the 129M/M mouse-passaged sCJD-VV2 prions were identical to those of the 129V/V mice inoculated with sCJD-VV2 prions. Moreover, we generated for the first time a type 2 PrPSc-specific antibody in addition to type 1 PrPSc-specific antibody and discovered that drastic changes in the PrPSc subpopulation underlie the traceback phenomenon. Here, we report the first direct evidence of the traceback in prion infection.Creutzfeldt-Jakob disease (CJD) is a lethal transmissible neurodegenerative disease caused by an abnormal isoform of prion protein (PrPSc), which is converted from the normal cellular isoform (PrPC) (1, 23). The genotype (M/M, M/V, or V/V, where M and V are allelotypes) at polymorphic codon 129 of the human prion protein (PrP) gene and the type (type 1 or type 2) of PrPSc in the brain are major determinants of the clinicopathological phenotypes of sporadic CJD (sCJD) (15-18). Type 1 and type 2 PrPSc are distinguishable according to the size of the proteinase K-resistant core of PrPSc (PrPres) (21 and 19 kDa, respectively), reflecting differences in the proteinase K cleavage site (at residues 82 and 97, respectively) (15, 18). According to this molecular typing system, sCJD can be classified into six subgroups (MM1, MM2, MV1, MV2, VV1, or VV2).The homology of the PrP genes between inoculated animals and the inoculum determines the susceptibility to prion infection. Transmission of sCJD prions to mice expressing human PrP with a nonhomologous genotype (referred to as cross-sequence transmission) results in a relatively long incubation period (10, 12). Meanwhile, the cross-sequence transmission can generate a new prion strain. Transmission of sCJD-VV2 prions to mice expressing human PrP with the 129M/M genotype generates unusual PrPres intermediate in size between type 1 and type 2 (10). We have designated this unusual PrPres with an upward size shift (Sh+) from the inoculated type 2 template MM[VV2]2Sh+ PrPres, where the notation is of the following form: host genotype [type of inoculated prion] type of generated PrPres.Similar to the MM[VV2]2Sh+ PrPres, the intermediate-sized PrPres has been observed in the plaque-type of dura mater graft-associated CJD (p-dCJD) (10, 13). Furthermore, a transmission study using p-dCJD prions revealed that PrP-humanized mice with the 129V/V genotype were highly susceptible to p-dCJD prions despite cross-sequence transmission (10). In addition, these 129V/V mice inoculated with p-dCJD prions produced type 2 PrPres (10). These findings suggest that p-dCJD could be caused by cross-sequence transmission of sCJD-VV2 prions to individuals with the 129M/M genotype. We have designated this phenomenon “traceback.” The traceback phenomenon was discovered for the first time by a transmission study using variant CJD (vCJD) prions (2). Mice expressing bovine PrP were highly susceptible to vCJD prions because vCJD was caused by cross-sequence transmission of bovine spongiform encephalopathy prions to human. These findings suggest that a traceback study can be a powerful tool to identify the origin of prions (2, 10, 11). However, the traceback phenomenon has not been verified experimentally despite the abundant circumstantial evidence described above.To verify the traceback of sCJD-VV2 prions, we inoculated sCJD-VV2 prions into PrP-humanized mice with the 129M/M genotype as an experimental model of p-dCJD. Thereafter, we inoculated these MM[VV2]2Sh+ prions into PrP-humanized mice with the 129M/M or 129V/V genotype and compared the incubation period, neuropathology, and the type of PrPres in the brain. Here, we report the first direct evidence of the traceback in prion infection.  相似文献   

10.
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   

11.
Prion diseases are a group of transmissible, invariably fatal neurodegenerative diseases that affect both humans and animals. According to the protein-only hypothesis, the infectious agent is a prion (proteinaceous infectious particle) that is composed primarily of PrPSc, the disease-associated isoform of the cellular prion protein, PrP. PrPSc arises from the conformational change of the normal, glycosylphosphatidylinositol (GPI)-anchored protein, PrPC. The mechanism by which this process occurs, however, remains enigmatic. Rabbits are one of a small number of mammalian species reported to be resistant to prion infection. Sequence analysis of rabbit PrP revealed that its C-terminal amino acids differ from those of PrP from other mammals and may affect the anchoring of rabbit PrP through its GPI anchor. Using a cell culture model, this study investigated the effect of the rabbit PrP-specific C-terminal amino acids on the addition of the GPI anchor to PrPC, PrPC localization, and PrPSc formation. The incorporation of rabbit-specific C-terminal PrP residues into mouse PrP did not affect the addition of a GPI anchor or the localization of PrP. However, these residues did inhibit PrPSc formation, suggesting that these rabbit-specific residues interfere with a C-terminal PrPSc interaction site.Prion diseases, traditionally known as transmissible spongiform encephalopathies (TSE), are a group of invariably fatal neurodegenerative diseases that affect both humans and animals. According to the protein-only hypothesis, an abnormal isoform of the host-encoded prion protein (PrPC), referred to as PrPSc, is the sole or major component of the infectious agent causing these diseases (33). These disorders affect a wide range of mammals and include diseases such as Creutzfeldt-Jakob disease (CJD), variant CJD, Gerstmann-Straüssler-Scheinker (GSS) syndrome, kuru, and fatal familial insomnia (FFI) in humans, scrapie in sheep and goats, chronic wasting disease (CWD) in cervids, and bovine spongiform encephalopathy (BSE) in cattle. The term “prion” was first used to describe the unique infectious agent and was derived from “proteinaceous infectious particle” to distinguish it from conventional pathogens such as bacteria and viruses (33).To date, rabbits are one of the few mammalian species reported to be resistant to prion infection. Rabbits do not develop clinical disease after inoculation with brain tissue from individuals affected by the human prion diseases CJD and kuru, or by a number of animal forms of the disease, including scrapie and transmissible mink encephalopathy (TME) (12). In addition, mouse neuroblastoma (MNB) cells overexpressing rabbit PrP are also resistant to prion infection (45). Evidence that rabbit cells per se have the correct cellular machinery to support prion propagation has come from studies using the rabbit kidney epithelial cell line RK13. Upon transfection with appropriate PrP-expressing transgenes, these cells are a highly efficient and robust model of prion infection (6, 25, 41, 43). RK13 cells do not have detectable levels of endogenous rabbit PrPC and are therefore ideal for studying exogenous PrPC and the propagation of prions from different species (6). Originally, it was shown that RK13 cells overexpressing ovine PrP became susceptible to infection with scrapie (43), and more recently, RK13 cells expressing rodent PrPC, from either the mouse or the bank vole, were readily infected by prions adapted to and propagated in these two species (6, 41). RK13 cells expressing human PrPC, however, were resistant to infection with human prions derived directly from a patient with sporadic CJD (25). Since RK13 cells overexpressing PrP are a well-established model of prion propagation, we can therefore conclude that while these cells apparently have the appropriate cellular machinery to support prion propagation, it may be a characteristic of the rabbit prion protein itself that results in the resistance of this species to prion infection. However, the loss of a cellular cofactor may also be a contributing factor.Analysis of the rabbit PrP amino acid sequence shows that it has all the features previously described for members of the PrP protein family, including an N-terminal signal peptide, an octapeptide repeat region, and a C-terminal signal sequence (26). While amino acid sequence comparison of both mouse and rabbit PrP species reveals 87% sequence homology, there are 22 amino acid differences between the two, and several of these reside in regions of PrP known to be important in PrPSc formation. In scrapie-infected MNB cells, the residues Gly99 and Met108 within the N terminus, Ser173 within the central region, and Ile214 within the C terminus of rabbit PrP were shown to inhibit PrPSc generation when incorporated into mouse PrP, suggesting that multiple amino acid residues in rabbit PrP inhibit PrPSc formation (45). Approximately one-third (9/33 residues in the immature sequence) of the amino acid difference between mouse and rabbit PrPs was shown to occur at the glycosylphosphatidylinositol (GPI) anchor attachment site (see Fig. S1 in the supplemental material). As yet, studies involving this region of rabbit PrP have not been performed. Therefore, this region of rabbit PrP may provide further insight into the resistance of rabbits to prion infection.GPI anchor addition occurs via a transamination reaction in the endoplasmic reticulum (ER) following cleavage of the C-terminal signal sequence (39). There is no consensus sequence with which to identify the C-terminal cleavage site, but there are three key C-terminal elements: (i) the cleavage site, or ω site, where the GPI anchor attaches to the COOH group of the ω amino acid; (ii) a hydrophilic spacer region of 8 to 12 amino acids (ω + 1 up to ω + 10); and (iii) a hydrophobic region of 10 to 20 amino acids (ω + 11 onwards) (9). Analysis of known GPI-anchored proteins has given rise to sequence motifs in the C-terminal signal peptide allowing the prediction of the ω site of proteins. Due to the complexity of experimentally determining the ω site of GPI-anchored proteins, relatively few of the many known GPI-anchored proteins have had their ω sites determined (36 of 340 proteins in 2008) (32) The ω site of hamster PrP was determined experimentally to be at amino acid 231 (34) and is predicted to be at the same site for PrPs from all mammals, based on amino acid sequence comparison. Amino acid substitutions near the ω site of mouse PrP revealed that mouse PrP has an ω site at residue 230 (17). It was also shown that single amino acid substitutions at and near the ω site of mouse PrP affect the anchoring and conversion efficiency of PrP (17). It is therefore possible that the amino acids at the C terminus and within the GPI anchor signal sequence of rabbit PrP lead to the resistance to prion infection.To date, no protein structures containing a GPI anchor have been determined by X-ray crystallography, and although the nuclear magnetic resonance (NMR) structures of mouse and rabbit PrP have been solved, they do not contain any structural information for the residues immediately preceding the GPI anchor. We therefore created a mutant mouse PrP model containing rabbit PrP-specific amino acids at the ω site to investigate whether these residues are involved in rabbit resistance to prion infection. Here we demonstrate that the GPI anchor attachment site is an important site that controls the ability of PrP to be converted into PrPSc and that residues ω and ω + 1 of PrP are important modulators of this pathogenic process.  相似文献   

12.
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   

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.
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   

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

18.
19.
Mouse bioassay remains the gold standard for determining proof of infectivity, strain type, and infectious titer estimation in prion disease research. The development of an approach using ex vivo cell-based assays remains an attractive alternative, both in order to reduce the use of mice and to hasten results. The main limitation of a cell-based approach is the scarcity of cell lines permissive to infection with natural transmissible spongiform encephalopathy strains. This study combines two advances in this area, namely, the standard scrapie cell assay (SSCA) and the Rov9 and MovS6 cell lines, which both express the ovine PrP VRQ allele, to assess to what extent natural and experimental ovine scrapie can be detected ex vivo. Despite the Rov9 and MovS6 cell lines being of different biological origin, they were both permissive and resistant to infection with the same isolates of natural sheep scrapie as detected by SSCA. Rov9 subclones that are 20 times more sensitive than Rov9 to SSBP/1-like scrapie infection were isolated, but all the subclones maintained their resistance to isolates that failed to transmit to the parental line. The most sensitive subclone of the Rov9 cell line was used to estimate the infectious titer of a scrapie brain pool (RBP1) and proved to be more sensitive than the mouse bioassay using wild-type mice. Increasing the sensitivity of the Rov9 cell line to SSBP/1 infection did not correlate with broadening susceptibility, as the specificity of permissiveness and resistance to other scrapie isolates was maintained.Prion diseases are a group of neurodegenerative diseases affecting humans and animals, including scrapie in sheep and goats and bovine spongiform encephalopathy (BSE) in cattle. A feature of prion diseases and, in particular, of scrapie, is the existence of different strains (6) which influence pathology and is most probably related to the conformation of the pathogenic form of the prion protein (PrPSc). The susceptibility of sheep to scrapie is determined by the PrP genotype; codons 136, 154, and 171 determine relative resistance and susceptibility, with amino acids valine (V), arginine (R), and glutamine (Q) at these positions (known as VRQ) being considered the sheep PrP allele most susceptible to classical scrapie (3).An array of diagnostic tests exist for prion diseases, aimed at the detection of the disease-associated protease-resistant form of the naturally occurring PrPC protein, termed PrPSc or PrPres after partial protease digestion. However, the level of detectable PrPSc does not quantitatively correlate with prion infectivity (2) and the current biochemical analysis of PrPSc cannot always determine the strain (6, 7).Mouse bioassay remains the gold standard for determining proof of infectivity, strain type, and infectious titer estimate in ruminant transmissible spongiform encephalopathy (TSE) research. Conventional mouse bioassays using wild-type mice are generally slow (>150 days, and considerably longer, >600, days for obtaining infectious titer information) and require multiple mice to be dosed (typically 6 or more) at each dilution of infectious material. Therefore, the development of an approach using ex vivo cell-based assays remains an ethically and economically desirable alternative. Using cell lines permissive to mouse-passaged scrapie strains, Klöhn et al. have developed a cell-based assay for measuring de novo infection and the titer of mouse-passaged scrapie (18).The main limitation of adopting a cell-based approach is the scarcity of cell lines permissive to infection with natural TSE strains (for a review, see references 31 and 34), as the majority of permissive cell lines can only be infected with rodent-adapted strains of scrapie and BSE (4, 9, 16, 20, 23, 24, 29, 33, 36). While there are currently no cell lines reported to be permissive to bovine BSE or human TSE diseases, there are cell lines which express ovine PrP that have been shown to be permissive to natural scrapie infection (1, 35). There is also one fibroblast-like deer cell line that is able to propagate chronic wasting disease (27).Two of the sheep scrapie-susceptible cell lines are the MovS6 cell line (1), a Schwann cell line derived from the tg301 transgenic mouse, and the Rov9 cell line (35), based on a stably transfected rabbit kidney epithelial cell line (RK13) that does not express endogenous PrP. Both express the VRQ allele of ovine PrP, the latter upon induction with doxycycline (35). These cell lines were found to be permissive to infection with a PrP genotype-matched VRQ homozygous scrapie field case, and de novo PrPSc maintained its phenotype when used as an inoculum in mouse bioassays (1, 35). Using fluorescence-activated cell sorting, Falanga et al. isolated Rov9 subclones that produce higher levels of PrPC and PrPSc than the parental cell line when infected (11).The primary objective of this study was to assess the permissiveness of the Rov9 and MovS6 cell lines to a panel of scrapie isolates from a range of sheep breeds with a range of PrP genotypes. Second, subcloning of the Rov9 cell line was undertaken in an attempt to identify subclones with greater sensitivity and more diverse permissibility to ovine scrapie isolates.  相似文献   

20.
Prion strain interference can influence the emergence of a dominant strain from a mixture; however, the mechanisms underlying prion strain interference are poorly understood. In our model of strain interference, inoculation of the sciatic nerve with the drowsy (DY) strain of the transmissible mink encephalopathy (TME) agent prior to superinfection with the hyper (HY) strain of TME can completely block HY TME from causing disease. We show here that the deposition of PrPSc, in the absence of neuronal loss or spongiform change, in the central nervous system corresponds with the ability of DY TME to block HY TME infection. This suggests that DY TME agent-induced damage is not responsible for strain interference but rather prions compete for a cellular resource. We show that protein misfolding cyclic amplification (PMCA) of DY and HY TME maintains the strain-specific properties of PrPSc and replicates infectious agent and that DY TME can interfere, or completely block, the emergence of HY TME. DY PrPSc does not convert all of the available PrPC to PrPSc in PMCA, suggesting the mechanism of prion strain interference is due to the sequestering of PrPC and/or other cellular components required for prion conversion. The emergence of HY TME in PMCA was controlled by the initial ratio of the TME agents. A higher ratio of DY to HY TME agent is required for complete blockage of HY TME in PMCA compared to several previous in vivo studies, suggesting that HY TME persists in animals coinfected with the two strains. This was confirmed by PMCA detection of HY PrPSc in animals where DY TME had completely blocked HY TME from causing disease.Prions are infectious agents of animals, including humans, which are comprised of PrPSc, a misfolded isoform of the noninfectious host encoded protein PrPC (17, 24, 50, 63). Prion diseases of humans are unique neurodegenerative disorders in that they can have either a sporadic, familial, or infectious etiology. Prions cause disease in economically important domestic and wild animal species such as bovine spongiform encephalopathy in cattle and chronic wasting disease in wild and captive cervids (20, 62). Prion diseases can be zoonotic as illustrated by the transmission of bovine spongiform encephalopathy to humans that resulted in the emergence of variant Creutzfeldt-Jacob disease (14, 19, 22, 23, 46, 61, 68). Prion diseases are inevitably fatal and there are currently no effective treatments (21).Prion strains are defined by a characteristic set of features that breed true upon experimental passage (33, 34). Strain-specific differences have been identified in incubation period, clinical signs, agent distribution, overdominance, host range, neuropathology, and biochemical properties of PrPSc (5, 10, 11, 13, 28, 34, 42, 44). Strain-specific conformations of PrPSc are hypothesized to encode prion strain diversity; however, it is not understood how these differences result in the distinct strain properties (11, 19, 40, 47, 59, 66).Prion strain interference may be involved in the emergence of a dominant strain from a mixture as could occur during prion adaptation to a new host species or during prion evolution (4, 36, 43, 48, 56). In the natural prion diseases, there are examples where an individual host may be infected with more than one prion strain (15, 25, 55, 57, 58). Experimentally, coinfection or superinfection of prion strains can result in interference where a blocking, long incubation period strain extends the incubation period or completely blocks a superinfecting, short incubation period strain from causing disease (26, 27). Prion interference has been described in experimental studies of mice and hamsters infected with a wide variety of prion strains and routes of inoculation, suggesting it may be a common property of prion disease (3, 27, 52, 53, 60).It has been proposed that prion strains compete for a shared “replication site”; however, mechanistic details are not known, and it is unclear whether the blocking strain destroys or occupies the replication sites required for the superinfecting strain (28). The transport to and relative onset of replication of interfering strains in a common population of neurons is an important factor that can determine which strain will emerge (8). In the present study, we sought to determine whether the blocking strain disables transport and spread of the superinfecting strain or whether prion interference is due to competition for a cellular resource.  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