首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
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   

2.
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   

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

4.
To understand the role of a crowded physiological environment in the pathogenesis of neurodegenerative diseases, we report the following. 1) The formation of fibrous aggregates of the human Tau fragment Tau-(244–441), when hyperphosphorylated by glycogen synthase kinase-3β, is dramatically facilitated by the addition of crowding agents. 2) Fibril formation of nonphosphorylated Tau-(244–441) is only promoted moderately by macromolecular crowding. 3) Macromolecular crowding dramatically accelerates amyloid formation by human prion protein. A sigmoidal equation has been used to fit these kinetic data, including published data of human α-synuclein, yielding lag times and apparent rate constants for the growth of fibrils for these amyloidogenic proteins. These biochemical data indicate that crowded cell-like environments significantly accelerate the nucleation step of fibril formation of human Tau fragment/human prion protein/human α-synuclein (a significant decrease in the lag time). These results can in principle be predicted based on some known data concerning protein concentration effects on fibril formation both in vitro and in vivo. Furthermore, macromolecular crowding causes human prion protein to form short fibrils and nonfibrillar particles with lower conformational stability and higher protease resistance activity, compared with those formed in dilute solutions. Our data demonstrate that a crowded physiological environment could play an important role in the pathogenesis of neurodegenerative diseases by accelerating amyloidogenic protein misfolding and inducing human prion fibril fragmentation, which is considered to be an essential step in prion replication.Amyloid fibrils associated with neurodegenerative diseases such as Alzheimer disease, Parkinson disease, Huntington disease, and transmissible spongiform encephalopathy (TSE)3 (15) can be considered biologically relevant failures of the cellular protein quality control mechanisms (6) consisting of molecular chaperones and proteases (7). Up to now, about 20 different proteins with unrelated sequences and tertiary structures are known to form fibrous aggregates associated with various neurodegenerative diseases. These amyloidogenic proteins include both natively unfolded proteins, such as human Tau protein (3) and human α-synuclein (8), and folded globular proteins such as human prion protein (4). There are two faces of protein misfolding in neurodegeneration as follows: a gain of toxic function and a loss of physiological function, which can even occur in combination (9).Human Tau protein, a marker for Alzheimer disease, forms filaments in the brains of patients with Alzheimer disease (3, 10, 11). It has been found that hyperphosphorylation of Tau reduces the binding affinity between Tau and tubulin and contributes to the self-association of Tau and the formation of Tau paired helical filaments (3, 1113). It has been proposed that glycogen synthase kinase-3β (GSK-3β) hyperphosphorylation of Tau plays an important role in Alzheimer disease (14, 15), and GSK-3β induces an Alzheimer disease-like hyperphosphorylation of Tau when overexpressed in cultured human neurons (16).A large body of data strongly suggests Creutzfeldt-Jakob disease, bovine spongiform encephalopathy, and other TSEs are caused by prions (4). Prions are infectious proteins that can transmit biological information by propagating protein misfolding and aggregation (17). The infectious agent is believed to consist entirely of the prion protein (PrP) and is devoid of nucleic acid (4, 17). Prion biogenesis is associated with the normal protease-sensitive form of the protein (cellular PrP molecule, PrPC) undergoing structural change into an abnormal, protease-resistant, disease-causing isoform of prion protein (PrPSc) (4, 17). Although the mechanism by which PrPC is converted to PrPSc in TSE-infected cells and in vivo is not clear, data from cell-free reactions suggest this process is akin to autocatalytic polymerization (18).Misfolding of Tau and prion proteins has been traditionally and widely studied in dilute solutions (10, 1921). However, the physiological environment is poorly modeled by such dilute solutions, and biochemical reactions in vivo differ greatly from those in dilute solutions (22). The proteins associated with neurodegenerative diseases form fibrils in a physiological environment crowded with other background macromolecules (2226), such as proteins, glycosaminoglycans, and proteoglycans (23). Crowding is not confined to cellular interiors but also occurs in the extracellular matrix of tissues (24) and takes place at membrane surfaces (27). For example, blood plasma contains ∼80 g/liter protein, a concentration sufficient to cause significant crowding effects (24). Polysaccharides also contribute to crowding, especially in the extracellular matrix of tissues such as collagen (23, 26). The conversion of PrP from a normal soluble conformation PrPC to its pathogenic conformation PrPSc is believed to occur on the cell surface, in the endocytic vesicles, or in the crowded extracellular matrix (18). Thus, macromolecular crowding on the cell surface and in the extracellular matrix may play an important role in the conformational transition and amyloid formation of PrP in vivo, which have not been fully characterized yet. In vitro, such a crowded environment can be achieved experimentally by adding high concentrations of single or mixed nonspecific crowding agents to the system (2331). Recently, it has been demonstrated that macromolecular crowding significantly enhances the rate of amyloid formation of α-synuclein (32, 33), amyloid-β peptides (27), and human apolipoprotein C-II (34). However, the role of the crowded physiological environment in the pathogenesis of neurodegenerative diseases is poorly understood so far.To address the contributions of crowded physiological environments on the pathogenesis of neurodegenerative diseases, we report here that macromolecular crowding dramatically accelerates fibril formation by human Tau fragment and by human prion protein under physiological conditions. Our results indicate that macromolecular crowding significantly accelerates the nucleation step of fibril formation of human Tau fragment/human prion protein/human α-synuclein by fitting the data to a sigmoidal equation (35, 36). Furthermore, macromolecular crowding causes human prion protein to form short fibrils and nonfibrillar particles with lower conformational stability and higher protease resistance activity, compared with those formed in dilute solutions.  相似文献   

5.
Cellular prion proteins, PrPC, carrying the amino acid substitutions P102L, P105L, or A117V, which confer increased susceptibility to human transmissible spongiform encephalopathies, are known to form structures that include transmembrane polypeptide segments. Herein, we investigated the interactions between dodecylphosphocholine micelles and the polypeptide fragments 90–231 of the recombinant mouse PrP variants carrying the amino acid replacements P102L, P105L, A117V, A113V/A115V/A118V, K110I/H111I, M129V, P105L/M129V, and A117V/M129V. Wild-type mPrP-(90–231) and mPrP[M129V]-(91–231) showed only weak interactions with dodecylphosphocholine micelles in aqueous solution at pH 7.0, whereas discrete interaction sites within the polypeptide segment 102–127 were identified for all other aforementioned mPrP variants by NMR chemical shift mapping. These model studies thus provide evidence that amino acid substitutions within the polypeptide segment 102–127 affect the interactions of PrPC with membranous structures, which might in turn modulate the physiological function of the protein in health and disease.Transmissible spongiform encephalopathies (TSEs),2 such as Creutzfeldt-Jakob disease and the Gerstmann-Sträussler-Scheinker syndrome in humans, are accompanied by the appearance in the brain of an aggregated “scrapie” isoform of the host-encoded prion protein, PrPSc (13). The cellular form, PrPC, consists of an unstructured N-terminal “tail” of residues 23–125 and a globular domain of residues 126–231, and is attached by a C-terminal glycosylphosphatidylinositol (GPI) anchor to the outer plasma membrane. This structure ensures a role of membrane interactions in the physiological function of PrPC and probably also in the disease-related events leading to TSEs. For example, transgenic mice expressing a prion protein variant lacking the GPI membrane anchor did not develop the typical clinical signs of TSE after inoculation with infectious brain homogenate, although significant amounts of PrPSc accumulated in the brain (4). This finding led to the conclusion that membrane-association of PrPC is necessary for the development of a TSE. Independent evidence for the importance of membrane interactions for the onset of prion diseases was derived from cell-free conversion assays and cell culture experiments (5, 6).Data have also been presented that indicate that in addition to the normal form with the C terminus linked to a GPI anchor and the C-terminal domain located on the cell surface, PrPC can adopt two different transmembrane topologies, CtmPrP and NtmPrP, which have the C-terminal polypeptide segment located in the lumen of the endoplasmic reticulum (CtmPrP) or in the cytoplasm (NtmPrP) (79). The population of the CtmPrP variant is <10% of the total wild-type prion protein present during cellular biosynthesis but is increased to 20–30% for the pathogenic mutations P102L, P105L, and A117V of human PrP and the designed variant mouse PrPs obtained with the amino acid exchanges A113V/A115V/A118V and K110I/H111I (1013). The population of CtmPrP was further increased when an additional mutation, L9R, was present in the N-terminal signal sequence (14), so that ∼50% of the PrP was synthesized as the CtmPrP variant in granule neurons obtained from transgenic mice expressing a prion protein construct carrying the four amino acid replacements L9R, A113V, A115V, and A118V (15). Quite generally, an increase in the population of CtmPrP was also shown to be associated with severe neurodegeneration in transgenic mice, and it has been suggested that CtmPrP may be the proximate cause of neuronal death in certain prion disorders (10, 11, 15).In vitro studies on interactions of full-length and N-terminally truncated forms of recombinant PrP showed that acidic membranes caused the N-terminal part of the protein to become more structured, whereas the C-terminal domain was destabilized (1619). Furthermore, zwitterionic gel-phase dipalmitoylphosphatidylcholine or raft-like membranes were shown to induce increased α-helical structure in recombinant Syrian hamster PrP-(90–231) at pH 7.0 (18, 19). Membrane interactions of polypeptides representing sequence motifs found in the prion protein have also been studied (2023).In this report we describe investigations of PrP interactions with a membrane mimetic and focus on the mutations P102L, P105L, and A117V, which have been linked with familial Gerstmann-Sträussler-Scheinker syndrome in humans (2, 24, 25). Our interest in these variant proteins is related to open questions about the mechanisms by which pathogenic mutations predispose humans for prion diseases. We studied the interactions of a recombinant wild-type mouse prion protein fragment, mPrP-(90–231), and the variants mPrP[P102L]-(91–231), mPrP[P105L]-(91–231), mPrP[A117V]-(90–231), mPrP[A113V,A115V,A118V]-(90–231), and mPrP[K110I,H111I]-(90–231). For these studies, we used the N-terminally truncated protein composed of residues 90–231. This region contains the transmembrane segment, all known disease-associated point mutations, the entire polypeptide fragment with proteinase K-resistance in PrPSc, which is also sufficient to transmit disease (1, 25, 26). The amino acid substitutions in these variant PrPs are located either within a hydrophobic stretch of residues 112–127, which is highly conserved in mammalian PrPs (27, 28), or in the positively charged segment of residues 95–111 (Fig. 1, B and C). We also included the M129V polymorphism into this study, which was reported to have a significant influence on the susceptibility of humans to prion diseases and on the disease phenotype. For example, the mutations P105L and A117V are only pathogenic in the presence of valine at position 129 (2, 24). The zwitterionic detergent dodecylphosphocholine (DPC, Fig. 1A) was used as a biomembrane mimetic model system, and NMR spectroscopy was employed to screen for protein-detergent micelle interactions, and for the structural characterization of the various prion protein constructs interacting with the detergent micelles.Open in a separate windowFIGURE 1.Detergent and proteins used in this study. A, zwitterionic form of DPC. B, schematic diagram of the mPrP-(90–231) polypeptide indicating the locations of the regular secondary structures, i.e. three α-helices and two strands of an antiparallel β-sheet, a “positively charged cluster” (CC) of amino acid residues in positions 95–111, and a “hydrophobic polypeptide segment” (HPS) comprising residues 112–127. C, amino acid sequence alignment of residues 90–135 for wild-type mPrP-(90–231) and the protein variants studied in this paper, where for each variant mPrP the amino acid replacements are given and identical residues are indicated by dots; the numbering is according to Schätzl et al. (27).  相似文献   

6.
Insulin plays a central role in the regulation of vertebrate metabolism. The hormone, the post-translational product of a single-chain precursor, is a globular protein containing two chains, A (21 residues) and B (30 residues). Recent advances in human genetics have identified dominant mutations in the insulin gene causing permanent neonatal-onset DM2 (14). The mutations are predicted to block folding of the precursor in the ER of pancreatic β-cells. Although expression of the wild-type allele would in other circumstances be sufficient to maintain homeostasis, studies of a corresponding mouse model (57) suggest that the misfolded variant perturbs wild-type biosynthesis (8, 9). Impaired β-cell secretion is associated with ER stress, distorted organelle architecture, and cell death (10). These findings have renewed interest in insulin biosynthesis (1113) and the structural basis of disulfide pairing (1419). Protein evolution is constrained not only by structure and function but also by susceptibility to toxic misfolding.Insulin plays a central role in the regulation of vertebrate metabolism. The hormone, the post-translational product of a single-chain precursor, is a globular protein containing two chains, A (21 residues) and B (30 residues). Recent advances in human genetics have identified dominant mutations in the insulin gene causing permanent neonatal-onset DM2 (14). The mutations are predicted to block folding of the precursor in the ER of pancreatic β-cells. Although expression of the wild-type allele would in other circumstances be sufficient to maintain homeostasis, studies of a corresponding mouse model (57) suggest that the misfolded variant perturbs wild-type biosynthesis (8, 9). Impaired β-cell secretion is associated with ER stress, distorted organelle architecture, and cell death (10). These findings have renewed interest in insulin biosynthesis (1113) and the structural basis of disulfide pairing (1419). Protein evolution is constrained not only by structure and function but also by susceptibility to toxic misfolding.  相似文献   

7.
A decoding algorithm is tested that mechanistically models the progressive alignments that arise as the mRNA moves past the rRNA tail during translation elongation. Each of these alignments provides an opportunity for hybridization between the single-stranded, -terminal nucleotides of the 16S rRNA and the spatially accessible window of mRNA sequence, from which a free energy value can be calculated. Using this algorithm we show that a periodic, energetic pattern of frequency 1/3 is revealed. This periodic signal exists in the majority of coding regions of eubacterial genes, but not in the non-coding regions encoding the 16S and 23S rRNAs. Signal analysis reveals that the population of coding regions of each bacterial species has a mean phase that is correlated in a statistically significant way with species () content. These results suggest that the periodic signal could function as a synchronization signal for the maintenance of reading frame and that codon usage provides a mechanism for manipulation of signal phase.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32]  相似文献   

8.
9.
Early onset generalized dystonia (DYT1) is an autosomal dominant neurological disorder caused by deletion of a single glutamate residue (torsinA ΔE) in the C-terminal region of the AAA+ (ATPases associated with a variety of cellular activities) protein torsinA. The pathogenic mechanism by which torsinA ΔE mutation leads to dystonia remains unknown. Here we report the identification and characterization of a 628-amino acid novel protein, printor, that interacts with torsinA. Printor co-distributes with torsinA in multiple brain regions and co-localizes with torsinA in the endoplasmic reticulum. Interestingly, printor selectively binds to the ATP-free form but not to the ATP-bound form of torsinA, supporting a role for printor as a cofactor rather than a substrate of torsinA. The interaction of printor with torsinA is completely abolished by the dystonia-associated torsinA ΔE mutation. Our findings suggest that printor is a new component of the DYT1 pathogenic pathway and provide a potential molecular target for therapeutic intervention in dystonia.Early onset generalized torsion dystonia (DYT1) is the most common and severe form of hereditary dystonia, a movement disorder characterized by involuntary movements and sustained muscle spasms (1). This autosomal dominant disease has childhood onset and its dystonic symptoms are thought to result from neuronal dysfunction rather than neurodegeneration (2, 3). Most DYT1 cases are caused by deletion of a single glutamate residue at positions 302 or 303 (torsinA ΔE) of the 332-amino acid protein torsinA (4). In addition, a different torsinA mutation that deletes amino acids Phe323–Tyr328 (torsinA Δ323–328) was identified in a single family with dystonia (5), although the pathogenic significance of this torsinA mutation is unclear because these patients contain a concomitant mutation in another dystonia-related protein, ϵ-sarcoglycan (6). Recently, genetic association studies have implicated polymorphisms in the torsinA gene as a genetic risk factor in the development of adult-onset idiopathic dystonia (7, 8).TorsinA contains an N-terminal endoplasmic reticulum (ER)3 signal sequence and a 20-amino acid hydrophobic region followed by a conserved AAA+ (ATPases associated with a variety of cellular activities) domain (9, 10). Because members of the AAA+ family are known to facilitate conformational changes in target proteins (11, 12), it has been proposed that torsinA may function as a molecular chaperone (13, 14). TorsinA is widely expressed in brain and multiple other tissues (15) and is primarily associated with the ER and nuclear envelope (NE) compartments in cells (1620). TorsinA is believed to mainly reside in the lumen of the ER and NE (1719) and has been shown to bind lamina-associated polypeptide 1 (LAP1) (21), lumenal domain-like LAP1 (LULL1) (21), and nesprins (22). In addition, recent evidence indicates that a significant pool of torsinA exhibits a topology in which the AAA+ domain faces the cytoplasm (20). In support of this topology, torsinA is found in the cytoplasm, neuronal processes, and synaptic terminals (2, 3, 15, 2326) and has been shown to bind cytosolic proteins snapin (27) and kinesin light chain 1 (20). TorsinA has been proposed to play a role in several cellular processes, including dopaminergic neurotransmission (2831), NE organization and dynamics (17, 22, 32), and protein trafficking (27, 33). However, the precise biological function of torsinA and its regulation remain unknown.To gain insights into torsinA function, we performed yeast two-hybrid screens to search for torsinA-interacting proteins in the brain. We report here the isolation and characterization of a novel protein named printor (protein interactor of torsinA) that interacts selectively with wild-type (WT) torsinA but not the dystonia-associated torsinA ΔE mutant. Our data suggest that printor may serve as a cofactor of torsinA and provide a new molecular target for understanding and treating dystonia.  相似文献   

10.
A Boolean network is a model used to study the interactions between different genes in genetic regulatory networks. In this paper, we present several algorithms using gene ordering and feedback vertex sets to identify singleton attractors and small attractors in Boolean networks. We analyze the average case time complexities of some of the proposed algorithms. For instance, it is shown that the outdegree-based ordering algorithm for finding singleton attractors works in time for , which is much faster than the naive time algorithm, where is the number of genes and is the maximum indegree. We performed extensive computational experiments on these algorithms, which resulted in good agreement with theoretical results. In contrast, we give a simple and complete proof for showing that finding an attractor with the shortest period is NP-hard.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32]  相似文献   

11.
12.
Collapsin response mediator protein 2 (CRMP2) is an intracellular protein that mediates signaling of Semaphorin3A (Sema3A), a repulsive axon guidance molecule. Fyn, a Src-type tyrosine kinase, is involved in the Sema3A signaling. However, the relationship between CRMP2 and Fyn in this signaling pathway is still unknown. In our research, we demonstrated that Fyn phosphorylated CRMP2 at Tyr32 residues in HEK293T cells. Immunohistochemical analysis using a phospho-specific antibody at Tyr32 of CRMP showed that Tyr32-phosphorylated CRMP was abundant in the nervous system, including dorsal root ganglion neurons, the molecular and Purkinje cell layer of adult cerebellum, and hippocampal fimbria. Overexpression of a nonphosphorylated mutant (Tyr32 to Phe32) of CRMP2 in dorsal root ganglion neurons interfered with Sema3A-induced growth cone collapse response. These results suggest that Fyn-dependent phosphorylation of CRMP2 at Tyr32 is involved in Sema3A signaling.Collapsin response mediator proteins (CRMPs)4 have been identified as intracellular proteins that mediate Semaphorin3A (Sema3A) signaling in the nervous system (1). CRMP2 is one of the five members of the CRMP family. CRMPs also mediate signal transduction of NT3, Ephrin, and Reelin (24). CRMPs interact with several intracellular molecules, including tubulin, Numb, kinesin1, and Sra1 (58). CRMPs are involved in axon guidance, axonal elongation, cell migration, synapse maturation, and the generation of neuronal polarity (1, 2, 4, 5).CRMP family proteins are known to be the major phosphoproteins in the developing brain (1, 9). CRMP2 is phosphorylated by several Ser/Thr kinases, such as Rho kinase, cyclin-dependent kinase 5 (Cdk5), and glycogen synthase kinase 3β (GSK3β) (2, 1013). The phosphorylation sites of CRMP2 by these kinases are clustered in the C terminus and have already been identified. Rho kinase phosphorylates CRMP2 at Thr555 (10). Cdk5 phosphorylates CRMP2 at Ser522, and this phosphorylation is essential for sequential phosphorylations by GSK3β at Ser518, Thr514, and Thr509 (2, 1113). These phosphorylations disrupt the interaction of CRMP2 with tubulin or Numb (2, 3, 13). The sequential phosphorylation of CRMP2 by Cdk5 and GSK3β is an essential step in Sema3A signaling (11, 13). Furthermore, the neurofibrillary tangles in the brains of people with Alzheimer disease contain hyperphosphorylated CRMP2 at Thr509, Ser518, and Ser522 (14, 15).CRMPs are also substrates of several tyrosine kinases. The phosphorylation of CRMP2 by Fes/Fps and Fer has been shown to be involved in Sema3A signaling (16, 17). Phosphorylation of CRMP2 at Tyr479 by a Src family tyrosine kinase Yes regulates CXCL12-induced T lymphocyte migration (18). We reported previously that Fyn is involved in Sema3A signaling (19). Fyn associates with PlexinA2, one of the components of the Sema3A receptor complex. Fyn also activates Cdk5 through the phosphorylation at Tyr15 of Cdk5 (19). In dorsal root ganglion (DRG) neurons from fyn-deficient mice, Sema3A-induced growth cone collapse response is attenuated compared with control mice (19). Furthermore, we recently found that Fyn phosphorylates CRMP1 and that this phosphorylation is involved in Reelin signaling (4). Although it has been shown that CRMP2 is involved in Sema3A signaling (1, 11, 13), the relationship between Fyn and CRMP2 in Sema3A signaling and the tyrosine phosphorylation site(s) of CRMPs remain unknown.Here, we show that Fyn phosphorylates CRMP2 at Tyr32. Using a phospho-specific antibody against Tyr32, we determined that the residue is phosphorylated in vivo. A nonphosphorylated mutant CRMP2Y32F inhibits Sema3A-induced growth cone collapse. These results indicate that tyrosine phosphorylation by Fyn at Tyr32 is involved in Sema3A signaling.  相似文献   

13.
SLC26A7 (human)/Slc26a7 (mouse) is a recently identified chloride-base exchanger and/or chloride transporter that is expressed on the basolateral membrane of acid-secreting cells in the renal outer medullary collecting duct (OMCD) and in gastric parietal cells. Here, we show that mice with genetic deletion of Slc26a7 expression develop distal renal tubular acidosis, as manifested by metabolic acidosis and alkaline urine pH. In the kidney, basolateral Cl/HCO3 exchange activity in acid-secreting intercalated cells in the OMCD was significantly decreased in hypertonic medium (a normal milieu for the medulla) but was reduced only mildly in isotonic medium. Changing from a hypertonic to isotonic medium (relative hypotonicity) decreased the membrane abundance of Slc26a7 in kidney cells in vivo and in vitro. In the stomach, stimulated acid secretion was significantly impaired in isolated gastric mucosa and in the intact organ. We propose that SLC26A7 dysfunction should be investigated as a potential cause of unexplained distal renal tubular acidosis or decreased gastric acid secretion in humans.The collecting duct segment of the distal kidney nephron plays a major role in systemic acid base homeostasis by acid secretion and bicarbonate absorption. The acid secretion occurs via H+-ATPase and H-K-ATPase into the lumen and bicarbonate is absorbed via basolateral Cl/HCO3 exchangers (14). The tubules, which are located within the outer medullary region of the kidney collecting duct (OMCD),2 have the highest rate of acid secretion among the distal tubule segments and are therefore essential to the maintenance of acid base balance (2).The gastric parietal cell is the site of generation of acid and bicarbonate through the action of cytosolic carbonic anhydrase II (5, 6). The intracellular acid is secreted into the lumen via gastric H-K-ATPase, which works in conjunction with a chloride channel and a K+ recycling pathway (710). The intracellular bicarbonate is transported to the blood via basolateral Cl/HCO3 exchangers (1114).SLC26 (human)/Slc26 (mouse) isoforms are members of a conserved family of anion transporters that display tissue-specific patterns of expression in epithelial cells (1524). Several SLC26 members can function as chloride/bicarbonate exchangers. These include SLC26A3 (DRA), SLC26A4 (pendrin), SLC26A6 (PAT1 or CFEX), SLC26A7, and SLC26A9 (2531). SLC26A7 and SLC26A9 can also function as chloride channels (3234).SLC26A7/Slc26a7 is predominantly expressed in the kidney and stomach (28, 29). In the kidney, Slc26a7 co-localizes with AE1, a well-known Cl/HCO3 exchanger, on the basolateral membrane of (acid-secreting) A-intercalated cells in OMCD cells (29, 35, 36) (supplemental Fig. 1). In the stomach, Slc26a7 co-localizes with AE2, a major Cl/HCO3 exchanger, on the basolateral membrane of acid secreting parietal cells (28). To address the physiological function of Slc26a7 in the intact mouse, we have generated Slc26a7 ko mice. We report here that Slc26a7 ko mice exhibit distal renal tubular acidosis and impaired gastric acidification in the absence of morphological abnormalities in kidney or stomach.  相似文献   

14.
15.
16.
Human concentrative nucleoside transporter 3 (hCNT3) utilizes electrochemical gradients of both Na+ and H+ to accumulate pyrimidine and purine nucleosides within cells. We have employed radioisotope flux and electrophysiological techniques in combination with site-directed mutagenesis and heterologous expression in Xenopus oocytes to identify two conserved pore-lining glutamate residues (Glu-343 and Glu-519) with essential roles in hCNT3 Na+/nucleoside and H+/nucleoside cotransport. Mutation of Glu-343 and Glu-519 to aspartate, glutamine, and cysteine severely compromised hCNT3 transport function, and changes included altered nucleoside and cation activation kinetics (all mutants), loss or impairment of H+ dependence (all mutants), shift in Na+:nucleoside stoichiometry from 2:1 to 1:1 (E519C), complete loss of catalytic activity (E519Q) and, similar to the corresponding mutant in Na+-specific hCNT1, uncoupled Na+ currents (E343Q). Consistent with close-proximity integration of cation/solute-binding sites within a common cation/permeant translocation pore, mutation of Glu-343 and Glu-519 also altered hCNT3 nucleoside transport selectivity. Both residues were accessible to the external medium and inhibited by p-chloromercuribenzene sulfonate when converted to cysteine.Physiologic nucleosides and the majority of synthetic nucleoside analogs with antineoplastic and/or antiviral activity are hydrophilic molecules that require specialized plasma membrane nucleoside transporter (NT)3 proteins for transport into or out of cells (14). NT-mediated transport is required for nucleoside metabolism by salvage pathways and is a critical determinant of the pharmacologic actions of nucleoside drugs (36). By regulating adenosine availability to purinoreceptors, NTs also modulate a diverse array of physiological processes, including neurotransmission, immune responses, platelet aggregation, renal function, and coronary vasodilation (4, 6, 7). Two structurally unrelated NT families of integral membrane proteins exist in human and other mammalian cells and tissues as follows: the SLC28 concentrative nucleoside transporter (CNT) family and the SLC29 equilibrative nucleoside transporter (ENT) family (3, 4, 6, 8, 9). ENTs are normally present in most, possibly all, cell types (4, 6, 8). CNTs, in contrast, are found predominantly in intestinal and renal epithelia and other specialized cell types, where they have important roles in absorption, secretion, distribution, and elimination of nucleosides and nucleoside drugs (13, 5, 6, 9).The CNT protein family in humans is represented by three members, hCNT1, hCNT2, and hCNT3. Belonging to a CNT subfamily phylogenetically distinct from hCNT1/2, hCNT3 utilizes electrochemical gradients of both Na+ and H+ to accumulate a broad range of pyrimidine and purine nucleosides and nucleoside drugs within cells (10, 11). hCNT1 and hCNT2, in contrast, are Na+-specific and transport pyrimidine and purine nucleosides, respectively (1113). Together, hCNT1–3 account for the three major concentrative nucleoside transport processes of human and other mammalian cells. Nonmammalian members of the CNT protein family that have been characterized functionally include hfCNT, a second member of the CNT3 subfamily from the ancient marine prevertebrate the Pacific hagfish Eptatretus stouti (14), CeCNT3 from Caenorhabditis elegans (15), CaCNT from Candida albicans (16), and the bacterial nucleoside transporter NupC from Escherichia coli (17). hfCNT is Na+- but not H+-coupled, whereas CeCNT3, CaCNT, and NupC are exclusively H+-coupled. Na+:nucleoside coupling stoichiometries are 1:1 for hCNT1 and hCNT2 and 2:1 for hCNT3 and hfCNT3 (11, 14). H+:nucleoside coupling ratios for hCNT3 and CaCNT are 1:1 (11, 16).Although much progress has been made in molecular studies of ENT proteins (4, 6, 8), studies of structurally and functionally important regions and residues within the CNT protein family are still at an early stage. Topological investigations suggest that hCNT1–3 and other eukaryote CNT family members have a 13 (or possibly 15)-transmembrane helix (TM) architecture, and multiple alignments reveal strong sequence similarities within the C-terminal half of the proteins (18). Prokaryotic CNTs lack the first three TMs of their eukaryotic counterparts, and functional expression of N-terminally truncated human and rat CNT1 in Xenopus oocytes has established that these three TMs are not required for Na+-dependent uridine transport activity (18). Consistent with this finding, chimeric studies involving hCNT1 and hfCNT (14) and hCNT1 and hCNT3 (19) have demonstrated that residues involved in Na+- and H+-coupling reside in the C-terminal half of the protein. Present in this region of the transporter, but of unknown function, is a highly conserved (G/A)XKX3NEFVA(Y/M/F) motif common to all eukaryote and prokaryote CNTs.By virtue of their negative charge and consequent ability to interact directly with coupling cations and/or participate in cation-induced and other protein conformational transitions, glutamate and aspartate residues play key functional and structural roles in a broad spectrum of mammalian and bacterial cation-coupled transporters (2030). Little, however, is known about their role in CNTs. This study builds upon a recent mutagenesis study of conserved glutamate and aspartate residues in hCNT1 (31) to undertake a parallel in depth investigation of corresponding residues in hCNT3. By employing the multifunctional capability of hCNT3 as a template for these studies, this study provides novel mechanistic insights into the molecular mechanism(s) of CNT-mediated cation/nucleoside cotransport, including the role of the (G/A)XKX3NEFVA(Y/M/F) motif.  相似文献   

17.
18.
A variety of high-throughput methods have made it possible to generate detailed temporal expression data for a single gene or large numbers of genes. Common methods for analysis of these large data sets can be problematic. One challenge is the comparison of temporal expression data obtained from different growth conditions where the patterns of expression may be shifted in time. We propose the use of wavelet analysis to transform the data obtained under different growth conditions to permit comparison of expression patterns from experiments that have time shifts or delays. We demonstrate this approach using detailed temporal data for a single bacterial gene obtained under 72 different growth conditions. This general strategy can be applied in the analysis of data sets of thousands of genes under different conditions.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29]  相似文献   

19.
Mathematical tools developed in the context of Shannon information theory were used to analyze the meaning of the BLOSUM score, which was split into three components termed as the BLOSUM spectrum (or BLOSpectrum). These relate respectively to the sequence convergence (the stochastic similarity of the two protein sequences), to the background frequency divergence (typicality of the amino acid probability distribution in each sequence), and to the target frequency divergence (compliance of the amino acid variations between the two sequences to the protein model implicit in the BLOCKS database). This treatment sharpens the protein sequence comparison, providing a rationale for the biological significance of the obtained score, and helps to identify weakly related sequences. Moreover, the BLOSpectrum can guide the choice of the most appropriate scoring matrix, tailoring it to the evolutionary divergence associated with the two sequences, or indicate if a compositionally adjusted matrix could perform better.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29]  相似文献   

20.
Transmissible spongiform encephalopathy (TSE) infectivity naturally spreads from site of entry in the periphery to the central nervous system where pathological lesions are formed. Several routes and cells within the host have been identified as important for facilitating the infectious process. Expression of the glycoprotein cellular PrP (PrPC) is considered a key factor for replication of infectivity in the central nervous system (CNS) and its transport to the brain, and it has been suggested that the infectious agent propagates from cell to cell via a domino-like effect. However, precisely how this is achieved and what involvement the different glycoforms of PrP have in these processes remain to be determined. To address this issue, we have used our unique models of gene-targeted transgenic mice expressing different glycosylated forms of PrP. Two TSE strains were inoculated intraperitoneally into these mice to assess the contribution of diglycosylated, monoglycosylated, and unglycosylated PrP in spreading of infectivity to the brain. This study demonstrates that glycosylation of host PrP has a profound effect in determining the outcome of disease. Lack of diglycosylated PrP slowed or prevented disease onset after peripheral challenge, suggesting an important role for fully glycosylated PrP in either the replication of the infectious agent in the periphery or its transport to the CNS. Moreover, mice expressing unglycosylated PrP did not develop clinical disease, and mice expressing monoglycosylated PrP showed strikingly different neuropathologic features compared to those expressing diglycosylated PrP. This demonstrates that targeting in the brain following peripheral inoculation is profoundly influenced by the glycosylation status of host PrP.Transmissible spongiform encephalopathies (TSE) or prion diseases are a group of fatal neurodegenerative diseases which include Creutzfeldt-Jakob disease (CJD) in humans, scrapie in sheep and goats, bovine spongiform encephalopathies (BSE) in cattle, and chronic wasting disease (CWD) in deer and elk (30). These diseases can be sporadic, familial, or acquired by infection, and the common hallmark is a distinct pathology in the central nervous system (CNS) characterized by neuronal loss, spongiform degeneration, and gliosis (38, 46).Expression of the host-encoded cellular PrP (PrPC) is fundamental for the onset of disease since PrP-deficient mice are refractory to TSE infection (11, 31). PrPC is a glycoprotein with two consensus sites for attachment of N-linked glycans (at codons 180 and 196 in the mouse) which are variably occupied, producing di-, mono-, and unglycosylated PrP (43). The diversity in glycosylation, combined with the complexity of added sugars, results in a large number of glycosylated forms of PrP (41). A central event associated with TSE infection is the conformational conversion of PrPC into an abnormal protease-resistant form, PrPSc (39). PrPSc is deposited in brain and, in some but not all cases, in peripheral organs of individuals affected by TSE (21).Although the pathology associated with TSE is found in the brain, the periphery is the most natural route of acquiring infection. Evidence suggests that oral transmission via contaminated food is linked with transmission of BSE to humans, resulting in variant CJD (vCJD) (10, 47), and blood transfusion has been identified as a probable route of human-to-human transmission of vCJD (23, 27, 36). Moreover, parenteral administration of contaminated human tissue-derived therapeutics has been shown to facilitate iatrogenic spread of these diseases (8, 46). It is therefore important to understand the mechanisms that allow the infectious agent to propagate in the periphery and be transported to the CNS prior to the onset of neurodegeneration in the brain.Many studies have been conducted to understand routes of transmission (for a review see references 1 and 29). Lymphoid tissues such as the spleen have been shown to play a fundamental role in agent replication and propagation in the very early stages of disease. Indeed, studies of splenectomized and asplenic mice have shown the lymphoreticular system (LRS) to be an important site for TSE agent replication (14, 26). The periphery also appears to have a role in processing the infectious agent following intracerebral (i.c.) inoculation as PrPSc accumulates in the spleen shortly after inoculation and before accumulation of the abnormal protein in the brain (15, 17). Within the LRS, follicular dendritic cells (FDC) have been shown to be important for the uptake of infectivity and subsequent spreading toward the CNS (7, 28, 33, 35). Several studies have also suggested the peripheral nervous systems (PNS) as a potential route of infectivity to the brain, implicating the vagus and sciatic nerves in this process (5, 20, 25, 34).Expression of PrPC in the peripheral tissues appears to be an important prerequisite for the transport of infectivity to the CNS following peripheral routes of inoculation. Indeed, it has been proposed that a continuous chain of cells expressing PrPC is fundamental for TSE neuroinvasion (6, 40), with overexpression of endogenous PrP in the PNS greatly facilitating the spread of infectivity (19). Thus, host PrP appears to have a fundamental role in the uptake, transport, and replication of the infectious agent (6). Moreover, it has been suggested that the different PrPC glycoforms may influence the timing of neuroinvasion by directly influencing the interaction with the infectious agent (19). However, the mechanism by which the different glycoforms are involved in these processes remains to be determined.In order to investigate the role of PrPC glycosylation in TSE disease after peripheral infection with different TSE strains, we have used our inbred gene-targeted transgenic mice expressing different glycosylated forms of PrP. These mice expressed PrP with no sugars at the first (designated G1/G1 in homozygous mice) or the second glycosylation site (G2/G2) or both (G3/G3) under the control of the endogenous PrP promoter (13). We have previously shown that following intracerebral inoculation, all glycotypes are susceptible to infection with at least one TSE strain and that the type of PrP glycosylation in the host influenced the incubation period but not the distribution of pathological lesions in the brain (45). Here, we examine the influence of host PrP glycosylation on the peripheral acquisition of infection and demonstrate that, unlike the intracerebral route, mice without PrP glycosylation were resistant to disease and that the different glycoforms had a profound influence on not only the timing of disease but also the type and distribution of the PrPSc deposits in the brain.  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