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.
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   

2.
Cytoplasmic localization of the prion protein (PrP) has been observed in different species and cell types. We have investigated this poorly understood phenomenon by expressing fusion proteins of sheep prion protein and green fluorescent protein (GFPPrP) in N2a cells, with variable sequence context surrounding the start codon Met1. GFPPrP expressed with the wild-type sequence was transported normally through the secretory pathway to the cell surface with acquisition of N-glycan groups, but two N-terminal fragments of GFPPrP were detected intracellularly, starting in frame from Met17. When GFPPrP was expressed with a compromised Kozak sequence (GFPPrP*), dispersed intracellular fluorescence was observed. A similar switch from pericellular to intracellular PrP localization was seen when analogous constructs of sheep PrP, without inserted GFP, were expressed, showing that this phenomenon is not caused by the GFP tag. Western blotting revealed a reduction in glycosylated forms of GFPPrP*, whereas the N-terminal fragments starting from Met17 were still present. Formation of these N-terminal fragments was completely abolished when Met17 was replaced by Thr, indicating that leaky ribosomal scanning occurs for normal sheep PrP and that translation from Met17 is the cause of the aberrant cytoplasmic localization observed for a fraction of the protein. In contrast, the same phenomenon was not detected upon expression of similar constructs for mouse PrP. Analysis of samples from sheep brain allowed immunological detection of N-terminal PrP fragments, indicating that sheep PrP is subject to similar processing mechanisms in vivo.PrPC 2 is a cell surface glycoprotein with an essential role in the pathogenesis of transmissible neurodegenerative prion diseases (1, 2). According to the prion hypothesis, a misfolded, pathogenic form of the protein (PrPSc) is the sole constituent of transmissible prions (3, 4), but the molecular details and required environs for the misfolding are incompletely understood. As would be expected for a glycosylphosphatidylinositol-anchored protein with N-linked glycans, PrPC is observed at the outer leaflet of the plasma membrane, the end point of the secretory route. The half-time at the plasma membrane is fairly short, because the protein may undergo shedding or endocytic internalization (59). Thus, PrPC can be encountered throughout the secretory and endocytic routes and is also able to leave cells via exosomes derived from multivesicular endosomes (10). In agreement with this, studies of the subcellular distribution of PrPC in mammalian brain have identified localization to the outer cell membrane, in the Golgi apparatus, and in endosomal vesicles (11, 12). However, others have found that PrPC is not solely associated with membranes, but, in some subpopulations of neurons, is localized to the cytoplasm (13, 14). In line with the latter observations, transgenic mice expressing PrP carrying a C-terminal GFP tag demonstrated intense cytoplasmic fluorescence from a limited number (approximately 1%) of the neurons in certain brain areas, such as the hippocampus (15). Immunohistochemical detection of intracellular, possibly cytoplasmic, PrP has also been reported from large mononuclear cells in the gut wall of sheep (16) and from enteric neurons in mice (17). The recent observations of pronounced cytoplasmic aggregation of PrP in pancreatic β-cells of rats prone to development of diabetes mellitus provide a perplexing example of nonstandard PrP localization in non-neuronal cells (18).The flexibility observed in the subcellular localization of PrPC has been suggested to be a requirement for normal functions of the protein (14, 19, 20), but how cytoplasmic and nuclear variants arise has not been established. Cytoplasmic PrP could be a result of retro-translocation from the endoplasmic reticulum (ER), as part of an unfolded protein response (2123) or from attenuated ER import of PrP under conditions of lumenal stress in the ER (24, 25). The finding of intact ER-targeting signal sequences on cytoplasmic PrPs (25, 26) favors the latter mechanism, namely a reduced ER import of PrP, possibly caused by saturation of the ER translocation machinery or an overload of unfolded proteins within the ER. However, no signs of stress or pathology could be detected in neurons of wild-type mice expressing cytoplasmic PrP (14), which led to the suggestion that the cytoplasmic appearance of PrP could constitute a physiologically relevant, but minor, pathway for the protein.Forced cytoplasmic expression of PrP in transgenic mice (22) and in the nematode Caenorhabditis elegans (27) resulted in neurodegenerative disease, suggesting that toxic mislocalization of PrP could be part of the pathogenic mechanism in prion diseases (28). However, transgenic mice expressing cytoplasmic PrP, on a PrP-null background, developed cerebellar atrophy but were resistant to experimental prion infection (29), suggesting that cytoplasmic PrP is unlikely to serve as substrate for prion replication. Furthermore, data obtained from transgenic mice expressing an anchorless secretory PrP show that, although these mice accumulate PrP-containing amyloid plaques upon challenge with PrPSc, they fail to develop clinical prion disease (30). Thus, membrane-attached PrP appears to be a prerequisite for development of prion-derived neurodegeneration.In eukaryotes, ribosomes bind specifically to linear mRNAs carrying a 7-methylguanosine 5′-end cap and slide along the mRNA in the 5′ → 3′ direction until they encounter the first start codon (AUG), from which the protein translation starts exclusively. Therefore, eukaryotic mRNAs are generally monocistronic. However, deviations from this standard principle have been reported, in which protein translation is initiated at alternative start codons either up or downstream from the primary AUG. The best characterized mechanism is known as context-dependent leaky ribosomal scanning (LRS) (31). This cap-dependent mechanism is particularly operative when the optimal (5′-GCCRCCaugG-3′) sequence context surrounding the first AUG codon is compromised, most notably at positions R−3 (where R= purine, A or G, but optimally G) and G+4 (32, 33).In this work, we report that in a cell culture system, sheep PrP mRNA displays a tendency to allow alternative translation initiation through LRS. Met17 serves as an internal in-frame alternative start codon giving rise to PrP with a severely shortened ER-targeting peptide.Although the LRS mechanism is active in sheep PrP, it appears to occur much less in mouse PrP (34). The molecular explanation and possible pathophysiological relevance of these observations in relation to PrP function await further studies. Interestingly, during the review process of this paper, observations of cytoplasmic PrP similar to some of those described herein were reported for human and hamster PrP (35).  相似文献   

3.
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   

4.
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.
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   

7.
It is now well established that the conversion of the cellular prion protein, PrPC, into its anomalous conformer, PrPSc, is central to the onset of prion disease. However, both the mechanism of prion-related neurodegeneration and the physiologic role of PrPC are still unknown. The use of animal and cell models has suggested a number of putative functions for the protein, including cell signaling, adhesion, proliferation, and differentiation. Given that skeletal muscles express significant amounts of PrPC and have been related to PrPC pathophysiology, in the present study, we used skeletal muscles to analyze whether the protein plays a role in adult morphogenesis. We employed an in vivo paradigm that allowed us to compare the regeneration of acutely damaged hind-limb tibialis anterior muscles of mice expressing, or not expressing, PrPC. Using morphometric and biochemical parameters, we provide compelling evidence that the absence of PrPC significantly slows the regeneration process compared to wild-type muscles by attenuating the stress-activated p38 pathway, and the consequent exit from the cell cycle, of myogenic precursor cells. Demonstrating the specificity of this finding, restoring PrPC expression completely rescued the muscle phenotype evidenced in the absence of PrPC.The cellular prion protein (PrPC) is a glycoprotein, prominently expressed in the mammalian central nervous system (CNS) and lymphoreticular system, that is anchored to the cell external surface through a glycolipidic moiety. The bad reputation acquired by PrPC originates from the notion that an aberrant conformer of it (PrPSc) is the major component of the prion, the unconventional infectious particle that causes fatal neurodegenerative disorders, i.e., transmissible spongiform encephalopathies (TSE) or prion diseases (56). A wealth of evidence has suggested that the function of PrPC is beneficial to the cell, but currently, our detailed comprehension of its physiology remains poor. In this respect, the availability of knockout (KO) paradigms for PrPC has provided less crucial information than expected. Subtle phenotypes, e.g., mild neuropathologic, cognitive, and behavioral deficits, have been described in PrP-KO mice (17, 50), but these animals generally live a normal life span without displaying obvious developmental defects (8, 42). Importantly, the same holds true when the expression of PrPC is postnatally abrogated (40). The extensive search for PrPC''s raison d''être has ascribed to the protein a plethora of functions (for updated reviews, see references 1 and 35); among these, roles in cell adhesion, migration, and differentiation have been proposed whereby PrPC could act by modulating different cell-signaling pathways (63). In this framework, a variety of neuronal proteins have been hypothesized to interact with PrPC (reviewed in references 1 and 11), for example, cell adhesion molecules or extracellular matrix proteins, which could explain the capacity of PrPC to mediate the neuritogenesis and neuronal differentiation observed in several cell model systems (13, 22, 23, 27, 36, 59, 64).Although neurons are generally regarded as the model of choice for unraveling the function of PrPC, the expression of the protein in several other organs suggests that PrPC has a conserved role in different tissues. Thus, important insight into PrPC function may also be provided by the analysis of extraneural tissues. One such tissue is skeletal muscle, which has been shown to express PrPC at significant levels (43, 46) and has been found to upregulate PrPC levels under stress conditions (71). On the other hand, ablation of the PrP gene has been shown to directly affect skeletal muscles, for example, by enhancing oxidative damage (30) or by diminishing tolerance for physical exercise (51). Skeletal muscles have also been associated with prion pathology, as evidenced by the accumulation of PrPSc (or PrPSc-like forms) in the muscles of TSE-affected humans and animals (2, 3, 6, 21, 53, 67) and by transgenic-mouse models of some inherited TSEs (16). In addition, overexpression of wild-type (WT) PrPC (25, 68), or expression of TSE-associated mutants of the protein (16, 66), generates myopathic traits in transgenic mice.In light of these notions, and because intact muscle tissues are more amenable to in vivo manipulations than neural tissue, we set out to analyze the potential role of PrPC in tissue morphogenesis (38, 41, 46) using an in vivo skeletal-muscle paradigm from two congenic mouse lines expressing (WT) or not expressing (PrP-KO) PrPC. Importantly, to verify that the PrP-KO muscle phenotype was specifically dependent on the absence of PrPC, we used PrP-KO mice reconstituted with a PrP transgene (PrP-Tg). The applied protocol consisted of first characterizing the degeneration of the hind-limb tibialis anterior (TA) muscle and then evaluating the myogenic process from the response to inflammation to the full recovery of the muscle. By combining acute insult with adult age, this strategy also had the potential to bypass possible compensatory mechanisms that might mask PrP-KO phenotypes during embryogenesis and/or in adulthood under normal conditions (65).In this study, we provide evidence that, compared to animals expressing PrPC (WT and PrP-Tg), recovery from damage of adult skeletal muscles was significantly slower in PrP-KO mice. Analysis of the different stages of muscle regeneration allowed us to conclude that PrPC is one of the factors that govern the early phases of this process, in which the proliferation and differentiation of myogenic precursor cells take place.  相似文献   

8.
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   

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.
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   

11.
Prion neuroinvasion from peripheral tissues involves agent replication in the lymphoreticular system (LRS) prior to entry into the nervous system. This study investigated the role of the LRS in prion neuroinvasion from the oral and nasal mucosa in wild-type and immunodeficient mice and in hamsters infected with the HY and DY strains of the transmissible mink encephalopathy (TME) agent. Following inoculation at neural sites, all hosts were susceptible to prion disease and had evidence of prion infection in the brain, but infection of the LRS was found only in scrapie-infected wild-type mice and HY TME-infected hamsters. In the LRS replication-deficient models, prion neuroinvasion was not observed following intraperitoneal or oral inoculation. However, immunodeficient mice, which have impaired follicular dendritic cells, were susceptible to scrapie following intratongue and intranasal inoculation despite the absence of PrPSc in the tongue or the nasal cavity. For DY TME, hamsters were susceptible following intratongue but not intranasal inoculation and PrPSc was limited to nerve fibers of the tongue. These findings indicate that neuroinvasion from the tongue and nasal cavity can be independent of LRS infection but neuroinvasion was partially dependent on the strain of the prion agent and/or the host species. The paucity of PrPSc deposition in the oral and nasal mucosa from LRS replication-deficient hosts following neuroinvasion from these tissues suggests an infection of nerve fibers that is below the threshold of PrPSc detection and/or the transport of the prion agent along cranial nerves without agent replication.In natural and experimental prion infections originating in the periphery, prion agent replication in the lymphoreticular system (LRS) precedes agent entry and spread in the peripheral nervous system. In the LRS, follicular dendritic cells (FDCs) are the major target of prion infection, and blocking or reversing FDC maturation can prevent scrapie agent replication in the LRS (25, 26, 28, 30, 32). Other migrating cell populations may also influence the progression of experimental prion disease (27, 36). From the LRS, centripetal spread of the prion agent to the spinal cord or brain occurs by spread along nerve fibers of the peripheral nervous system. In the central nervous system, prion agent replication can induce neurodegeneration and disease after an incubation period that can last from weeks to years. For example, in lambs from flocks with endemic scrapie, agent replication is initially detected in the gut-associated lymphoid tissues prior to proximal and distal spread in the LRS, infection of peripheral nerves that innervate the LRS, and subsequent spread to the spinal cord (19, 42). In addition, scrapie agent infection of the vagal nerve, which innervates many peripheral organs including the digestive tract, results in axonal transport directly to the dorsal motor nucleus of the vagus in the brain stem (29, 41). The role of scrapie infection in the LRS in the latter pathway of neuroinvasion is unknown. A similar pathway of prion neuroinvasion occurs in mule deer experimentally infected with the chronic wasting disease agent with the exception that early infection is also established in the lymph nodes of the upper gastrointestinal tract (37, 38). Recent studies indicate that a similar pathway of neuroinvasion occurs in natural and experimental bovine spongiform encephalopathy (BSE) following oral exposure except that agent replication in the LRS is greatly reduced and appears restricted to portions of the gut-associated lymphoid tissues (13, 20, 39).There are natural prion diseases in sheep and cattle that do not exhibit the typical distribution of the prion agent in the brain and LRS that are presumably acquired via oral prion exposure (5, 33). The absence of the abnormal isoform of the prion protein, PrPSc, in the LRS and dorsal motor nucleus of the vagus in atypical scrapie and the H type or L type of BSE raises the question as to whether these cases are due to infection by an alternate route(s) other than ingestion or whether these cases have an etiology that is distinct from that of acquired prion diseases. Direct prion infection of nerve fibers or terminals in highly innervated tissues, such as the mucosa in the head, has been suggested to represent potential sites of prion agent entry that would not require prior agent replication in the LRS (4, 12, 31). The presence of scrapie or BSE infection in the retina, sensory fibers of the tongue, and nasal mucosa of sheep, goat, and/or cattle suggests that the eye, tongue, or nasal cavity could be alternate sites of prion agent entry into hosts (8, 11, 15, 16, 40). Experimental prion inoculation at these mucosal sites can cause prion disease and in some cases rapid neuroinvasion (4, 9, 17, 18). Another explanation for this distribution of infection is that centrifugal spread of the prion agent away from the brain and along cranial nerves could serve as a pathway for prion infection and accumulation in these mucosal tissues (4, 10, 43).In this work, we investigated the role of the LRS in prion neuroinvasion from the oral and nasal cavities. In order to investigate neuroinvasion following neural and extraneural routes of inoculation in which prion replication is blocked in the LRS, we used two rodent models for prion infection. In muMT mice, which lack mature B cells, and in lymphotoxin-α (LTα) null mice, FDCs do not undergo maturation, and as a result, these mice do not develop clinical disease following intraperitoneal inoculation of the scrapie agent but are susceptible following direct inoculation into the brain (23, 30). In a second model, the HY and DY strains of the transmissible mink encephalopathy (TME) agent were used to investigate neuroinvasion in Syrian hamsters. The HY and DY TME agents can replicate in the nervous system, but the DY TME agent does not replicate in the LRS, and therefore, the DY TME agent is not pathogenic following intraperitoneal (i.p.) inoculation (2, 3). Following intratongue (i.t.) or intranasal (i.n.) inoculation, prion neuroinvasion was independent of scrapie agent replication in the LRS of immunodeficient mice, but evidence for scrapie infection of peripheral nerve fibers or olfactory neurons at these mucosa was lacking. In hamsters, i.t. inoculation of the HY or DY TME agent resulted in PrPSc deposition in nerve fibers and prion disease, but only the HY TME agent caused disease following i.n. inoculation. These findings suggest that neuroinvasion from the oral and nasal mucosa in LRS replication-deficient rodents can be independent of LRS infection, but the paucity of PrPSc at these mucosal sites of exposure in immunodeficient mice and DY TME-infected hamsters suggests that neuroinvasion is due to either a low-level prion infection of the nervous system at the site of inoculation or transport of the prion agent in axons in the absence of agent replication at the site of prion entry. These findings indicate that these mucosal tissues may not exhibit early evidence of infection and therefore will prove difficult to identify as a portal for agent entry.  相似文献   

12.
Fibrils play an important role in the pathogenesis of amyloidosis; however, the underlying mechanisms of the growth process and the structural details of fibrils are poorly understood. Crucial in the fibril formation of prion proteins is the stacking of PrP monomers. We previously proposed that the structure of the prion protein fibril may be similar as a parallel left-handed β-helix. The β-helix is composed of spiraling rungs of parallel β-strands, and in the PrP model residues 105–143 of each PrP monomer can contribute two β-helical rungs to the growing fibril. Here we report data to support this model. We show that two cyclized human PrP peptides corresponding to residues 105–124 and 125–143, based on two single rungs of the left-handed β-helical core of the human PrPSc fibril, show spontaneous cooperative fibril growth in vitro by heterologous stacking. Because the structural model must have predictive value, peptides were designed based on the structure rules of the left-handed β-helical fold that could stack with prion protein peptides to stimulate or to block fibril growth. The stimulator peptide was designed as an optimal left-handed β-helical fold that can serve as a template for fibril growth initiation. The inhibiting peptide was designed to bind to the exposed rung but frustrate the propagation of the fibril growth. The single inhibitory peptide hardly shows inhibition, but the combination of the inhibitory with the stimulatory peptide showed complete inhibition of the fibril growth of peptide huPrP-(106–126). Moreover, the unique strategy based on stimulatory and inhibitory peptides seems a powerful new approach to study amyloidogenic fibril structures in general and could prove useful for the development of therapeutics.Transmissible spongiform encephalopathies are neurodegenerative disorders in a wide range of mammalian species, including Creutzfeldt-Jacob disease in man, scrapie in sheep, and bovine spongiform encephalopathy in cattle. The deposition of aggregated prion protein fibrils on and in neurons is regarded to be the source of these neurodegenerative diseases and is frequently associated with occurrence of Congo red positivity (13). The fibrils are formed by the conformational change of the prion protein (PrPc)2 into the scrapie form (PrPSc). The misfolded conformer of the prion protein (PrPSc) is considered as the causative agent in these diseases according to the protein-only hypothesis (4). Studies have shown the toxicity of fibrils of the full-length recombinant mammalian prion protein as well as soluble β-rich oligomers to cultured cells and primary neurons (5).It is still unknown how much of the whole PrPSc molecule is involved in the fibril growth. It is shown that the N-terminal part of PrP, specifically residues 112–141, can go through conformational changes involving β-strand formation, which subsequently triggers fibril growth (68), and solid state NMR studies showed that residues 112–141 are part of the highly ordered core of huPrP-(23–144) (9). It was previously shown that peptides based on the 89–143 region of the human PrP protein can form fibrils rich in β-sheet structure which are biologically active in transgenic mice (10). Within this region it is the huPrP-(106–126) peptide that is the smallest known region of PrP that forms fibrils that are toxic and resemble the physiological properties of PrPSc (1116). The formation of PrPSc is considered to be a two-step event; first, there is the binding between PrPc and PrPSc and subsequently the conformational conversion from PrPc into PrPSc occurs. Mutation studies in a prion-infected neuroblastoma cell line showed that in mouse PrP the regions 101–110 and 136–158 are crucial for the binding and conversion events, respectively (17). Because prevention of fibril growth is the prime therapeutic target, detailed structural knowledge of the fibril is essential for understanding the mechanism of fibril growth. However, structural analysis of amyloid fibrils is hampered by insolubility, isomorphism, and aggregation. X-ray diffraction of several amyloid fibrils revealed a so-called cross-β diffraction pattern which indicates that the fibrils contain β-strands perpendicular to the fibril axis and hydrogen bonds in parallel (18, 19). Thus, for fibril growth the β-strands have to stack on top of each other. Several structures have been suggested to explain the structure of the stacked β-strands; e.g. a parallel in register organization of stacked β hairpins (24) or the comparable dry steric zipper structure (25). Previously, we and other groups suggested that the β-sheet structures in the PrPSc fibril may be similar to the topologically most simple class of β-sheets; that is, the parallel left-handed β-helix (Fig. 1A) (6, 20, 21). The left-handed β helix is formed by triangular progressive coils (rungs) of 18–20 residues. Each rung is formed by three hexapeptide motifs, which results in an approximate 3-fold symmetry. Backbone-backbone hydrogen bonding and stacking of the side chains in adjacent rungs contribute to the folding of β-helical rungs. We suggested that each PrPSc monomer contributes two left-handed β-helical rungs to the fibril, comprising residues 105–124 and 125–143 (Fig. 1A). This two-rung structural model was recently confirmed for amyloid fibrils of the HET-s prion by NMR analysis (22). In contrast to fibrils which are composed of homologous stacks of identical peptides, e.g. the Aβ peptide (23), the PrPSc fibril is more complex because it is composed of heterologous stacks of at least two peptides. For homologous stacking of two identical peptides, the complementarity issue is relatively simple because the identical side chains are in register (e.g. Ile-Ile, Val-Val stacking, and Asn ladders). However, in the case of heterologous stacking, the side chains of the additional heterologous peptide needs to be complementary with the other peptide to allow fibril growth.Open in a separate windowFIGURE 1.A, theoretical model of the fibrillogenic core of PrPSc. In the PrPSc model based on the left-handed β-helix structure, each PrPSc monomer contributes two stacked rungs to the fibril (different color for each monomer). The protofibril is formed by consecutive stacking of the two windings. The stack of two rungs provides enough elevation to accommodate the remaining part (residues ∼ 146–253) of the PrPSc molecule (20). B, the left-handed β-helix structure of LpxA-based on x-ray crystallography. In the left-handed β-helix structure of LpxA (PDB code 1LXA) rungs 6 and 7 are indicated (red) that were used for the heterologous stacking studies. Linear and cyclized peptides based on rung 6 and rung 7 were modified to satisfy the ideal left-handed β-helix motif (see “LpxA Peptides” under “Results”) and tested for their intrinsic and cooperative fibrillogenicity. C, left-handed β-helical rung based on rung 6 of LpxA. The rung is formed by three hexapeptide motifs, which results in an approximate 3-fold symmetry. A left-handed β-helical rung can be cyclized by a disulfide bridge after the introduction of a cysteine at position 2 of the first hexapeptide and position 1 of the fourth hexapeptide (according to the numbering used for the hexapeptide repeats in the left-handed β-helix).To investigate whether the suggested rungs 105–123 and 125–143 from human PrP could be complementary (20), we studied the homologous stacking and the heterologous stacking of linear and cyclized prion protein peptides comprising the huPrP-(105–143) region (KTNMKHMAGAAAAGAVVGGLGGYMLGSAMSRPIIHFGS). Qualitative and semiquantitative analysis were done by electron microscopy and Congo red staining. The quantification of the fibril formation was assessed by thioflavin S staining, in which the addition of polyanions (e.g. heparin) enhance the β-sheet formation of peptides comprising the 82–143 region of PrP and improve the reproducibility of the fibril growth (24). This study provides first evidence of heterologous stacking by two isolated putative β-strand layers (or rungs) of the human prion protein with fibril formation as a result. The left-handed β-helix structure provided insight for the “stack-and-stop” approach. With this approach a mix of a stimulatory peptide and an inhibitory peptide could completely block fibril formation. The stimulatory peptide was based on the 125–143 region that was optimized to serve as a folding template for the consecutive stacking of the 106–126 peptide. This cooperative fibril growth was completely inhibited by the inhibitory peptide based on peptides 106–126 with strategic d-amino acid and/or proline substitutions. The findings in this study support models in which the sequential strands in a fibril must somehow spiral up- or downward along the fibril axis, e.g. like the hypothetical left-handed β-helical structure of PrPSc fibrils (20). Furthermore, it allows the development of well defined small protein modules which can be used for structure studies of the 82–143 domain of PrPSc and the development of therapeutics.  相似文献   

13.
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   

14.
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   

15.
16.
The binding of the adaptor protein APPL1 to adiponectin receptors is necessary for adiponectin-induced AMP-activated protein kinase (AMPK) activation in muscle, yet the underlying molecular mechanism remains unknown. Here we show that in muscle cells adiponectin and metformin induce AMPK activation by promoting APPL1-dependent LKB1 cytosolic translocation. APPL1 mediates adiponectin signaling by directly interacting with adiponectin receptors and enhances LKB1 cytosolic localization by anchoring this kinase in the cytosol. Adiponectin also activates another AMPK upstream kinase Ca2+/calmodulin-dependent protein kinase kinase by activating phospholipase C and subsequently inducing Ca2+ release from the endoplasmic reticulum, which plays a minor role in AMPK activation. Our results show that in muscle cells adiponectin is able to activate AMPK via two distinct mechanisms as follows: a major pathway (the APPL1/LKB1-dependent pathway) that promotes the cytosolic localization of LKB1 and a minor pathway (the phospholipase C/Ca2+/Ca2+/calmodulin-dependent protein kinase kinase-dependent pathway) that stimulates Ca2+ release from intracellular stores.Adiponectin, an adipokine abundantly expressed in adipose tissue, exhibits anti-diabetic, anti-inflammatory, and anti-atherogenic properties and hence is a potential therapeutic target for various metabolic diseases (13). The beneficial effects of adiponectin are mediated through the direct interaction of adiponectin with its cell surface receptors, AdipoR1 and AdipoR2 (4, 5). Adiponectin increases fatty acid oxidation and glucose uptake in muscle cells by activating AMP-activated protein kinase (AMPK)3 (4, 6), which depends on the interaction of AdipoR1 with the adaptor protein APPL1 (Adaptor protein containing Pleckstrin homology domain, Phosphotyrosine binding domain, and Leucine zipper motif) (5). However, the underlying mechanisms by which APPL1 mediates adiponectin signaling to AMPK activation and other downstream targets remain unclear.AMPK is a serine/threonine protein kinase that acts as a master sensor of cellular energy balance in mammalian cells by regulating glucose and lipid metabolism (7, 8). AMPK is composed of a catalytic α subunit and two noncatalytic regulatory subunits, β and γ. The NH2-terminal catalytic domain of the AMPKα subunit is highly conserved and contains the activating phosphorylation site (Thr172) (9). Two AMPK variants, α1 and α2, exist in mammalian cells that show different localization patterns. AMPKα1 subunit is localized in non-nuclear fractions, whereas the AMPKα2 subunit is found in both nucleus and non-nuclear fractions (10). Biochemical regulation of AMPK activation occurs through various mechanisms. An increase in AMP level stimulates the binding of AMP to the γ subunit, which induces a conformational change in the AMPK heterotrimer and results in AMPK activation (11). Studies have shown that the increase in AMPK activity is not solely via AMP-dependent conformational change, rather via phosphorylation by upstream kinases, LKB1 and CaMKK. Dephosphorylation by protein phosphatases is also important in regulating the activity of AMPK (12).LKB1 has been considered as a constitutively active serine/threonine protein kinase that is ubiquitously expressed in all tissues (13, 14). Under conditions of high cellular energy stress, LKB1 acts as the primary AMPK kinase through an AMP-dependent mechanism (1517). Under normal physiological conditions, LKB1 is predominantly localized in the nucleus. LKB1 is translocated to the cytosol, either by forming a heterotrimeric complex with Ste20-related adaptor protein (STRADα/β) and mouse protein 25 (MO25α/β) or by associating with an LKB1-interacting protein (LIP1), to exert its biological function (1822). Although LKB1 has been shown to mediate contraction- and adiponectin-induced activation of AMPK in muscle cells, the underlying molecular mechanisms remain elusive (15, 23).CaMKK is another upstream kinase of AMPK, which shows considerable sequence and structural homology with LKB1 (2426). The two isoforms of CaMKK, CaMKKα and CaMKKβ, encoded by two distinct genes, share ∼70% homology at the amino acid sequence level and exhibit a wide expression in rodent tissues, including skeletal muscle (2734). Unlike LKB1, AMPK phosphorylation mediated by CaMKKs is independent of AMP and is dependent only on Ca2+/calmodulin (35). Hence, it is possible that an LKB1-independent activation of AMPK by CaMKK exists in muscle cells. However, whether and how adiponectin stimulates this pathway in muscle cells are not known.In this study, we demonstrate that in muscle cells adiponectin induces an APPL1-dependent LKB1 translocation from the nucleus to the cytosol, leading to increased AMPK activation. Adiponectin also activates CaMKK by stimulating intracellular Ca2+ release via the PLC-dependent mechanism, which plays a minor role in activation of AMPK. Taken together, our results demonstrate that enhanced cytosolic localization of LKB1 and Ca2+-induced activation of CaMKK are the mechanisms underlying adiponectin-stimulated AMPK activation in muscle cells.  相似文献   

17.
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   

18.
Human FGF1 (fibroblast growth factor 1) is a powerful signaling molecule with a short half-life in vivo and a denaturation temperature close to physiological. Binding to heparin increases the stability of FGF1 and is believed to be important in the formation of FGF1·fibroblast growth factor receptor (FGFR) active complex. In order to reveal the function of heparin in FGF1·FGFR complex formation and signaling, we constructed several FGF1 variants with reduced affinity for heparin and with diverse stability. We determined their biophysical properties and biological activities as well as their ability to translocate across cellular membranes. Our study showed that increased thermodynamic stability of FGF1 nicely compensates for decreased binding of heparin in FGFR activation, induction of DNA synthesis, and cell proliferation. By stepwise introduction of stabilizing mutations into the K118E (K132E) FGF1 variant that shows reduced affinity for heparin and is inactive in stimulation of DNA synthesis, we were able to restore the full mitogenic activity of this mutant. Our results indicate that the main role of heparin in FGF-induced signaling is to protect this naturally unstable protein against heat and/or proteolytic degradation and that heparin is not essential for a direct FGF1-FGFR interaction and receptor activation.FGF1 (fibroblast growth factor 1) belongs to a family of polypeptide growth factors comprising in humans 22 structurally related proteins (1, 2). The signaling induced by the growth factor leads to a wide range of cellular responses during development as well as in adult life, such as growth regulation, differentiation, survival, stress response, migration, and proliferation of different cell types (3). The biological activity of FGF1 is exerted through binding to four high affinity cell surface receptors (FGFR1–4), resulting in receptor dimerization and transphosphorylation in its tyrosine kinase domain (4, 5). The activated FGFR3 induces cellular response by initiating several signaling cascades, including mitogen-activated protein kinase (MAPK), phosphoinositide 3-kinase/Akt, and phospholipase C-γ (PLC-γ) pathways (6).In addition to FGFRs, FGF1 binds to heparan sulfates (HS) associated with proteoglycans at the cell surface and in the extracellular matrix (7). Among the physiological sugars, the highest affinity for FGF1 is shown by heparin, a widely used linear, highly sulfated polysaccharide composed of 2-O-sulfated iduronic acid and 6-O-sulfated, N-sulfated glucosamine units (8).Despite many years of research, there is still controversy regarding the molecular role of heparin/HS in FGF1- and FGF2-induced signaling. Thus, the question of whether or not the linkage of two molecules of the growth factor by heparin/HS is an absolute prerequisite for induction of FGFR dimerization is still open. Numerous studies have concluded that the presence of heparin/HS is obligatory for FGF signaling. It is widely believed that heparin/HS is directly involved in receptor dimerization and is critical for mitogenic response stimulated by the growth factor (4, 6, 810).On the other hand, several authors working on FGF1 and FGF2 have suggested that there is no mandatory requirement for heparin for the assembly and activation of the FGF·FGFR complex. They imply that heparin only plays a role in association of two molecules of the growth factor and therefore facilitates their binding to FGFR (11). It has been reported that FGF1 and FGF2 can interact with the FGFR and trigger phosphorylation of p42/44 MAPK and activation of other signaling pathways even in the absence of HS (1216).The accepted role of heparin/HS in FGF1 signaling is to prevent the degradation of the growth factor (17). The interaction with heparin or HS protects FGF1 against heat, acidic pH, and proteases (18, 19). HS also seems to regulate the activity of different FGFs by creating their local reservoir and generating a concentration gradient of the growth factor (6, 17).The binding of FGF1 to heparin/HS is mediated by specific residues forming a positively charged patch on the protein surface (20, 21). The major contribution is made by Lys118 (Lys132 in the full-length numbering system), which was identified by Harper and Lobb (22), and Lys112 and Arg122 (23, 24). Additional residues of FGF1 involved in the interaction with heparin are the positively charged Lys113, Arg119, and Lys128 and the polar Asn18, Asn114, and Gln127 (20, 21). Site-directed mutagenesis and other studies have revealed the importance of Lys118 not only in heparin binding but also for the biological function of FGF1 (22, 25, 26). It was shown that the K118E (K132E) mutant is inactive in stimulation of DNA synthesis, although its affinity for FGFR and the ability to activate signaling cascades is not reduced (27, 28). Despite extensive research, the reason for the lack of mitogenic potential of K118E FGF1 is still not clear.In this paper, we verified the function of heparin in FGF1·FGFR complex formation and signaling by constructing several FGF1 mutants with reduced affinity for heparin. To recover the stability of these variants, which could no longer be stabilized by heparin, we supplemented them stepwise with stabilizing mutations (29). We analyzed thoroughly their biological activity and their ability to translocate across cellular membranes (3034). Interestingly, the full mitogenic activity of the K118E FGF1 variant was restored by the introduced stabilizing mutations.Our results indicate that the main role of heparin in FGF-induced signaling is to protect this naturally unstable protein against heat denaturation and proteolytic degradation and that the increased stability of the growth factor can compensate for reduced heparin binding.  相似文献   

19.
Age-related macular degeneration (AMD) causes severe vision loss in the elderly; early identification of AMD risk could help slow or prevent disease progression. Toward the discovery of AMD biomarkers, we quantified plasma protein Nε-carboxymethyllysine (CML) and pentosidine from 58 AMD and 32 control donors. CML and pentosidine are advanced glycation end products that are abundant in Bruch membrane, the extracellular matrix separating the retinal pigment epithelium from the blood-bearing choriocapillaris. We measured CML and pentosidine by LC-MS/MS and LC-fluorometry, respectively, and found higher mean levels of CML (∼54%) and pentosidine (∼64%) in AMD (p < 0.0001) relative to normal controls. Plasma protein fructosyl-lysine, a marker of early glycation, was found by amino acid analysis to be in equal amounts in control and non-diabetic AMD donors, supporting an association between AMD and increased levels of CML and pentosidine independent of other diseases like diabetes. Carboxyethylpyrrole (CEP), an oxidative modification from docosahexaenoate-containing lipids and also abundant in AMD Bruch membrane, was elevated ∼86% in the AMD cohort, but autoantibody titers to CEP, CML, and pentosidine were not significantly increased. Compellingly higher mean levels of CML and pentosidine were present in AMD plasma protein over a broad age range. Receiver operating curves indicate that CML, CEP adducts, and pentosidine alone discriminated between AMD and control subjects with 78, 79, and 88% accuracy, respectively, whereas CML in combination with pentosidine provided ∼89% accuracy, and CEP plus pentosidine provided ∼92% accuracy. Pentosidine levels appeared slightly altered in AMD patients with hypertension and cardiovascular disease, indicating further studies are warranted. Overall this study supports the potential utility of plasma protein CML and pentosidine as biomarkers for assessing AMD risk and susceptibility, particularly in combination with CEP adducts and with concurrent analyses of fructosyl-lysine to detect confounding factors.Age-related macular degeneration (AMD)1 is a progressive, multifactorial disease and a major cause of severe vision loss in the elderly (1). Deposition of debris (drusen) in the macular region of Bruch membrane, the extracellular matrix separating the choriocapillaris from the retinal pigment epithelium (RPE), is an early, hallmark risk factor of AMD. The disease can progress to advanced dry AMD (geographic atrophy), which is characterized by regional degeneration of photoreceptor and RPE cells, or to advanced wet AMD (choroidal neovascularization (CNV)), which is characterized by abnormal blood vessels growing from the choriocapillaris through Bruch membrane beneath the retina. CNV accounts for over 80% of debilitating vision loss in AMD; however, only 10–15% of AMD cases progress to CNV.There is growing consensus that AMD is an age-related inflammatory disease involving dysregulation of the complement system; however, triggers of the inflammatory response have yet to be well defined. Oxidative stress appears to be involved as smoking significantly increases the risk of AMD (2), antioxidant vitamins can selectively slow AMD progression (3), and a host of oxidative protein and DNA modifications have been detected at elevated levels in AMD Bruch membrane, drusen, retina, RPE, and plasma (411). Oxidative protein modifications like carboxyethylpyrrole (CEP) and Nε-carboxymethyllysine (CML), both elevated in AMD Bruch membrane, stimulate neovascularization in vivo (12, 13), suggesting possible roles in CNV. Other studies have shown that mice immunized with CEP protein modifications develop an AMD-like phenotype (14). Accordingly oxidative modifications may be catalysts or triggers of AMD pathology (6).AMD has long been hypothesized to be a systemic disease (15) based in part on the presence of retinal drusen in patients with membranoproliferative glomerulonephritis type II (16) and systemic complement activation in AMD (17). Support for this hypothesis also comes from mounting evidence that advanced glycation end products (AGEs) may play a role in AMD (4, 5, 7, 18, 19). AGEs are a heterogeneous group of mostly oxidative modifications resulting from the Maillard nonenzymatic glycation reaction that have been associated with age-related diseases and diabetic complications (20, 21). In 1998, CML was the first AGE to be found in AMD Bruch membrane and drusen (4). Other AGEs have since been detected in AMD ocular tissues (5, 7, 18) and in Bruch membrane, drusen, RPE, and choroidal extracellular matrix from healthy eyes (6, 22). CML, a nonfluorescent AGE, and pentosidine, a fluorescent cross-linking AGE, increase with age in Bruch membrane (18, 23). Receptors for AGEs (RAGE and AGE-R1) appear elevated on RPE and photoreceptor cells in early and advanced dry AMD (7) especially in RPE overlying drusen-like deposits on Bruch membrane (19). AGE-R3, also known as galectin-3, is elevated in AMD Bruch membrane (24).Although AMD susceptibility genes now account for over 50% of AMD cases (25), many individuals with AMD risk genotypes may never develop advanced disease with severe vision loss. Nevertheless the prevalence of advanced AMD is increasing (26). Toward the discovery of better methods to detect those at risk for advanced AMD, we quantified CML and pentosidine in plasma proteins from AMD and control patients and compared their discriminatory accuracy with plasma CEP biomarkers. CEP biomarkers have been shown to enhance the AMD predictive accuracy of genomic AMD biomarkers (11). This report shows CML and pentosidine to be elevated in AMD plasma proteins and demonstrates their potential biomarker utility in assessing AMD risk and susceptibility especially in combination with CEP biomarkers.  相似文献   

20.
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