首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
2.
3.
4.
5.
During propagation, yeast prions show a strict sequence preference that confers the specificity of prion assembly. Although propagations of [PSI+] and [RNQ+] are independent of each other, the appearance of [PSI+] is facilitated by the presence of [RNQ+]. To explain the [RNQ+] effect on the appearance of [PSI+], the cross-seeding model was suggested, in which Rnq1 aggregates act as imperfect templates for Sup35 aggregation. If cross-seeding events take place in the cytoplasm of yeast cells, the collision frequency between Rnq1 aggregates and Sup35 will affect the appearance of [PSI+]. In this study, to address whether cross-seeding occurs in vivo, a new [PSI+] induction method was developed that exploits a protein fusion between the prion domain of Sup35 (NM) and Rnq1. This fusion protein successfully joins preexisting Rnq1 aggregates, which should result in the localization of NM around the Rnq1 aggregates and hence in an increased collision frequency between NM and Rnq1 aggregates. The appearance of [PSI+] could be induced very efficiently, even with a low expression level of the fusion protein. This study supports the occurrence of in vivo cross-seeding between Sup35 and Rnq1 and provides a new tool that can be used to dissect the mechanism of the de novo appearance of prions.Prions were originally defined as proteinaceous infectious particles responsible for transmissible spongiform encephalopathies in mammals (reviewed in reference 23). Since a non-Mendelian genetic element, [URE3], was identified as a yeast prion (37), however, this concept has been expanded to include protein-based genetic elements. In addition to [URE3], there are at least two more proteinaceous genetic elements in Saccharomyces cerevisiae, namely, [PSI+] and [RNQ+] (20, 22, 28). [Het-s] was also identified as a prion in the filamentous fungus Podospora anserina (2).Despite the absence of any structural and functional homologies between various prion proteins from different organisms, they share a common feature, i.e., prion proteins can adopt two distinct conformational states. One of these, the aggregated prion state, can stimulate the soluble, nonprion conformation to convert into the prion form. Gaining intermolecular β-sheet structures, purified yeast prion proteins aggregate and form amyloid fibers in vitro (8, 12, 28, 32). Protein extract from yeast cells in the prion state can facilitate the in vitro polymerization of soluble prion protein from nonprion cells (21), and amyloid fibers of purified yeast prion proteins can convert the cells into the prion state when introduced into yeast cells, demonstrating the protein-only hypothesis (15, 31). Thus, intracellular prion aggregates are thought to have the same structural basis as amyloid fibers formed in vitro.Yeast prion biology has provided invaluable insights into the prion concept at the molecular level. Because of its experimental convenience, [PSI+] has been investigated most intensively among various yeast prions. [PSI+] results from the aggregation of Sup35 protein, which is essential for terminating the translation at stop codons. When Sup35 is in the [PSI+] aggregated state, ribosomes often fail to release polypeptides at stop codons, causing a non-Mendelian trait which is easily detected by nonsense suppression. ade1 or ade2 nonsense mutants are used as marker genes to determine the [PSI+] state. These mutants cannot grow on adenine-deficient medium and form red colonies on medium supplemented with a limiting amount of adenine, such as yeast extract-peptone-dextrose (YPD). ade mutants in the [PSI+] state, however, can grow on adenine-deficient medium and form white colonies, as they produce functional Ade1 or Ade2 by virtue of a nonsense mutation readthrough. To sustain propagation, all yeast prions need the disaggregation activity of Hsp104, which can be inhibited by guanidine hydrochloride (GuHCl) (9). Since yeast prions are cured by growth on guanidine-containing medium, prion phenotypes can easily be distinguished from chromosomal suppressor mutants.Sup35 (eRF3) of S. cerevisiae has a prion-determining N-terminal domain (N), a highly charged middle domain (M) that confers solubility on the molecule, and an essential C-terminal domain that binds guanine nucleotides and stimulates the polypeptide release reaction catalyzed by Sup45 (eRF1) (17, 29, 33). The de novo appearance of [PSI+] can be induced by overexpression of SUP35 or its prion domain-containing fragments (NM) (6). [PSI+] induction, however, can be achieved only in [RNQ+] cells that harbor the prion state of the Rnq1 protein (4, 19). Two hypotheses about how [RNQ+] can affect the appearance of [PSI+] have been suggested. One is an inhibitor titration model that postulates the molecules preventing the aggregation of Sup35 and the recruitment of these inhibitors to Rnq1 aggregates in [RNQ+] cells. The other is a cross-seeding model in which Rnq1 aggregates directly catalyze the polymerization of Sup35. In vitro cross-seeding between different amyloidogenic proteins was reported, and Rnq1 amyloid fiber can also act as a seed for Sup35 polymerization in vitro (7, 13). These in vitro data support the possibility of cross-seeding between Rnq1 and Sup35. However, because the milieu of cytoplasm is very different from that of a test tube, whether this cross-seeding really occurs in vivo is still obscure. For this study, we developed a new, robust [PSI+] induction method that confirms the cross-seeding events in the cytoplasmic environment.  相似文献   

6.
The prion hypothesis13 states that the prion and non-prion form of a protein differ only in their 3D conformation and that different strains of a prion differ by their 3D structure.4,5 Recent technical developments have enabled solid-state NMR to address the atomic-resolution structures of full-length prions, and a first comparative study of two of them, HET-s and Ure2p, in fibrillar form, has recently appeared as a pair of companion papers.6,7 Interestingly, the two structures are rather different: HET-s features an exceedingly well-ordered prion domain and a partially disordered globular domain. Ure2p in contrast features a very well ordered globular domain with a conserved fold, and—most probably—a partially ordered prion domain.6 For HET-s, the structure of the prion domain is characterized at atomic-resolution. For Ure2p, structure determination is under way, but the highly resolved spectra clearly show that information at atomic resolution should be achievable.Key words: prion, NMR, solid-state NMR, MAS, structure, Ure2p, HET-sDespite the large interest in the basic mechanisms of fibril formation and prion propagation, little is known about the molecular structure of prions at atomic resolution and the mechanism of propagation. Prions with related properties to the ones responsible for mammalian diseases were also discovered in yeast and funghi8,9 which provide convenient model system for their studies. Prion proteins described include the mammalian prion protein PrP, Ure2p,10 Rnq1p,11 Sup35,12 Swi1,13 and Cyc8,14 from bakers yeast (S. cervisiae) and HET-s from the filamentous fungus P. anserina. The soluble non-prion form of the proteins characterized in vitro is a globular protein with an unfolded, dynamically disordered N- or C-terminal tail.1518 In the prion form, the proteins form fibrillar aggregates, in which the tail adopts a different conformation and is thought to be the dominant structural element for fibril formation.Fibrills are difficult to structurally characterize at atomic resolution, as X-ray diffraction and liquid-state NMR cannot be applied because of the non-crystallinity and the mass of the fibrils. Solid-state NMR, in contrast, is nowadays well suited for this purpose. The size of the monomer, between 230 and 685 amino-acid residues for the prions of Figure 1, and therefore the number of resonances in the spectrum—that used to be large for structure determination—is now becoming tractable by this method.Open in a separate windowFigure 1Prions identified today and characterized as consisting of a prion domain (blue) and a globular domain (red).Prion proteins characterized so far were found to be usually constituted of two domains, namely the prion domain and the globular domain (see Fig. 1). This architecture suggests a divide-and-conquer approach to structure determination, in which the globular and prion domain are investigated separately. In isolation, the latter, or fragments thereof, were found to form β-sheet rich structures (e.g., Ure2p(1-89),6,19 Rnq1p(153-405)20 and HET-s(218-289)21). The same conclusion was reached by investigating Sup35(1-254).22 All these fragements have been characterized as amyloids, which we define in the sense that a significant part of the protein is involved in a cross-beta motif.23 An atomic resolution structure however is available presently only for the HET-s prion domain, and was obtained from solid-state NMR24 (vide infra). It contains mainly β-sheets, which form a triangular hydrophobic core. While this cross-beta structure can be classified as an amyloid, its triangular shape does deviate significantly from amyloid-like structures of smaller peptides.23Regarding the globular domains, structures have been determined by x-ray crystallography (Ure2p25,26 and HET-s27), as well as NMR (mammal prions15,2830). All reveal a protein fold rich in α-helices, and dimeric structures for the Ure2 and HET-s proteins. The Ure2p fold resembles that of the β-class glutathione S-transferases (GST), but lacks GST activity.25It is a central question for the structural biology of prions if the divide-and-conquer approach imposed by limitations in current structural approaches is valid. Or in other words: can the assembly of full-length prions simply be derived from the sum of the two folds observed for the isolated domains?  相似文献   

7.
Yeast prions are heritable protein-based genetic elements which rely on molecular chaperone proteins for stable transmission to cell progeny. Within the past few years, five new prions have been validated and 18 additional putative prions identified in Saccharomyces cerevisiae. The exploration of the physical and biological properties of these “nouveau prions” has begun to reveal the extent of prion diversity in yeast. We recently reported that one such prion, [SWI+], differs from the best studied, archetypal prion [PSI+] in several significant ways.1 Notably, [SWI+] is highly sensitive to alterations in Hsp70 system chaperone activity and is lost upon growth at elevated temperatures. In that report we briefly noted a correlation amongst prions regarding amino acid composition, seed number and sensitivity to the activity of the Hsp70 chaperone system. Here we extend that analysis and put forth the idea that [SWI+] may be representative of a class of asparagine-rich yeast prions which also includes [URE3], [MOT3+] and [ISP+], distinct from the glutamine-rich prions such as [PSI+] and [RNQ+]. While much work remains, it is apparent that our understanding of the extent of the diversity of prion characteristics is in its infancy.Key words: Sis1, Hsp40, chromatin remodeling, Swi1, Ssa, heat-shock, protein misfolding, cell stress, Hsp 104, PINYeast prions are heritable elements, most of which are amyloid aggregates of single proteins. The three best studied yeast prions [PSI+], [RNQ+] (also called [PIN+]), and [URE3] are formed from amyloid aggregates of the cytosolic yeast proteins Sup35, Rnq1 and Ure2, respectively.2 Yeast prions can spontaneously arise in an otherwise clonal cell population, a process referred to as prion formation or nucleation, but once formed their continued propagation is intimately related to molecular chaperone activity. Chaperone function is needed to fragment prion amyloids to create heritable seeds which can then be passed on to cell progeny, thus maintaining the prion in the cell line.3 Yeast prions vary in the steady-state number of heritable seeds per cell; having more seeds increases the chances of passing the prion to progeny and hence prions with higher seed numbers are more mitotically stable.46The currently accepted model of prion fragmentation posits that components of the Hsp70 chaperone system work in congress with the disaggregase Hsp104.1,79 Hsp70-type chaperones function by repeatedly binding and releasing client polypeptides in an ATP-dependent manner, a cycle that is tightly regulated by co-chaperone proteins (Fig. 1). J-proteins (Hsp40s) stimulate Hsp70 ATP hydrolysis and peptide binding via a conserved J-domain whereas nucleotide exchange factors (NEFs) stimulate ADP/ATP exchange, restoring the ATP-bound (peptide unbound) state. In prion fragmentation, the J-protein Sis1, the Hsp70 Ssa, and nucleotide exchange factors (NEFs) of the Sse family are co-chaperones required as partners for the Hsp70 Ssa. While chaperone proteins may have additional functions in prion biology, e.g., prion formation, these additional functions are still poorly understood.9Open in a separate windowFigure 1The Cyclic Hsp70 Chaperone System. Ssa (purple), the yeast cytosolic Hsp70, binds and releases client polypeptides (blue) in a regulated and ATP-dependent manner. J-proteins (aquamarine) including Sis1, Ydj1 and others, stimulate Ssa ATP hydrolysis by virtue of a conserved J-domain and thereby catalyze the “forward” direction of the cycle as indicated above. ADP•Ssa more stably associates with client polypeptides than the ATP-bound form and hence J-proteins favor the ADP•Ssa•Peptide complex. In some cases, J-proteins can also bind and deliver client polypeptides to Hsp70s via C-terminal domains (also shown above). Nucleotide exchange factors (NEFs), including the Sse proteins (dark blue) which share some structural homology with Ssa, catalyze the “reverse” direction of the cycle by facilitating ADP release and subsequent ATP binding, and thus favor an ATP•Ssa state with a dissociated peptide.In the past few years, the number of known yeast prions has rapidly grown such that, to date, a total of eight yeast prions have been identified and an additional 18 proteins have been annotated as putative prions.10 The biological and physical properties of these newly discovered prions are only beginning to be explored. We recently reported the results of an investigation into the biological properties of the prion [SWI+], which is formed from the chromatin-remodeling factor Swi1.1 Swi1 is part of the SWI/SNF chromatin-remodeling complex that regulates the expression of approximately 6% of all yeast genes.11 The presence of [SWI+] causes partial loss of SWI/SNF chromatin-remodeling function, resulting in the impaired ability to uptake certain sugars, among other phenotypes.11 [SWI+] is a prion of particular interest because of its potential to alter global gene expression. Below we describe its intriguing interactions with molecular chaperone proteins and environmental stress, and the implications of these properties on yeast prion biology.  相似文献   

8.
9.
Neurodegenerative diseases are caused by proteinaceous aggregates, usually consisting of misfolded proteins which are often typified by a high proportion of β-sheets that accumulate in the central nervous system. These diseases, including Morbus Alzheimer, Parkinson disease and Transmissible Spongiform Encephalopathies (TSEs)—also termed prion disorders—afflict a substantial proportion of the human population and, as such, the etiology and pathogenesis of these diseases has been the focus of mounting research. Although many of these diseases arise from genetic mutations or are sporadic in nature, the possible horizontal transmissibility of neurodegenerative diseases poses a great threat to population health. In this article we discuss recent studies that suggest that the “non-transmissible” status bestowed upon Alzheimer and Parkinson diseases may need to be revised as these diseases have been successfully induced through tissue transplants. Furthermore, we highlight the importance of investigating the “natural” mechanism of prion transmission including peroral and perenteral transmission, proposed routes of gastrointestinal uptake and neuroinvasion of ingested infectious prion proteins. We examine the multitude of factors which may influence oral transmissibility and discuss the zoonotic threats that Chronic Wasting disease (CWD), Bovine Spongiform Encephalopathy (BSE) and Scrapie may pose resulting in vCJD or related disorders. In addition, we suggest that the 37 kDa/67 kDa laminin receptor on the cell surface of enterocytes, a major cell population in the intestine, may play an important role in the intestinal pathophysiology of alimentary prion infections.Key words: prion, 37 kDa/67 kDa laminin receptor, CJD, BSE, CWD, scrapie, Alzheimer disease, Parkinson disease, intestine, enterocytesMany different mechanisms exist which underlie the etiology of the numerous neurodegenerative diseases affecting the human population. Amongst the most prominent are Morbus Alzheimer, prion disorders, Parkinson disease, Chorea Huntington, frontotemporal dementia and amylotrophic lateral sclerosis. The molecular mechanisms underlying these diseases vary; however, all neurodegenerative diseases share a common feature: they are caused by protein aggregation. The only neurodegenerative diseases proven to be transmissible are prion disorders. In contrast to frontotemporal dementia, recent evidence suggests that Alzheimer and Parkinson diseases may also be transmissible. Pre-symptomatic Alzheimer disease (APP23) mice exhibited an increase in the Alzheimer phenotype when brain homogenate of autopsied human Alzheimer disease patients and older, amyloid beta- (Aβ-) laden APP23 mice was injected into their hippocampi.1 These findings suggest that the Aβ-abundant brain homogenate of Alzheimer disease patients may possess the ability to induce or supplement the overproduction of Aβ, possibly leading to the onset of Alzheimer disease.The pathological feature associated with Parkinson disease is the formation of Lewy bodies in cell bodies and neuronal processes in the brain.2 The main component of these protein aggregates is α-synuclein (reviewed in ref. 2). Autopsies of Parkinson disease patients revealed that Lewy bodies had formed on healthy embryonic neurons that had been grafted onto the brain tissue of the patients several years before (prior to said examination).35 It may thus be proposed that α-synuclein transmission is possible from diseased to healthy neurons, suggesting that Parkinson disease may be transmissible from a Parkinson disease patient to a healthy individual. These findings imply that Alzheimer and Parkinson diseases may be transmissible through tissue transplants and the use of contaminated surgical tools.6Prion disorders, also termed Transmissible Spongiform Encephalopathies (TSEs), are fatal neurodegenerative diseases that affect the central nervous system (CNS) of multiple animal species. In lieu of the social, economic and political ramifications of such infections, as well as the possible intra- and interspecies transmissibility of such disorders, various routes of experimental transmission have been investigated including intracerebral, intraperitoneal, intraventricular, intraocular, intraspinal and subcutaneous injections (reviewed in ref. 79). However, such routes of transmission are not representative of the “natural” mechanism as the majority of prion disorders are contracted through ingestion of infectious prion (PrPSc) containing material. Thus, the peroral and perenteral prion transmission is of greatest consequence with respect to TSE disease establishment. Moreover, the presence of PrPSc in the buccal cavity of scrapie-infected sheep10 (reviewed in ref. 11) and the possible horizontal transfer as a result hereof, as may be similarly proposed for animals suffering from other TSEs, may further contribute to the oral transmissibility of TSEs.A number of model systems have been employed to study TSE transmissibility. Owing to ethical constraints, TSE transmissibility to humans via the oral route may not be directly investigated and as a result hereof, alternative model systems are needed. These may include the use of transgenic mice, cell lines which are permissive to infection12 and experimental animals such as sheep, calves, goats, minks, ferrets and non-human primates (reviewed in ref. 9).Intestinal entry of PrPSc has been proposed to occur via two pathways, the membranous (M) cell-dependent and M cell-independent pathways (Fig. 1).13,14 The former involves endocytic M (microfold)-cells, which cover the intestinal lymphoid follicles (Peyer''s patches)14 and may take up prions and thereby facilitate the translocation of these proteins across the intestinal epithelium into the lymphoid tissues (reviewed in ref. 9) as has been demonstrated in a cellular model.13 Following such uptake by the M cells, the prions may subsequently pass to the dendritic cells and follicular dendritic cells (FDCs) (Fig. 1), which allow for prion transport to the mesenteric lymph nodes and replication, respectively.15 The prion proteins may subsequently gain access to the enteric nervous system (ENS) and ultimately the central nervous system (CNS).15Open in a separate windowFigure 1Proposed routes of gastrointestinal entry of ingested infectious prions (PrPSc) as well as possible pathways of amplification and transport to the central nervous system.However, prion intestinal translocation has been observed in the absence of M cells and has been demonstrated to be as a result of the action of polar, 37 kDa/67 kDa LRP/LR (non-integrin laminin receptor; reviewed in ref. 1618) expressing enterocytes. Enterocytes are the major cell population of the intestinal epithelium and due to their ability to endocytose pathogens, nutrients and macromolecules,19 it has been proposed that these cells may represent a major entry site for alimentary prions (Fig. 1).Since enterocyte prion uptake has been demonstrated to be dependent on the presence of LRP/LR on the apical brush border of the cells,14,20 the interaction between varying prion protein strains and the receptor2123 may be employed as a model system to study possible oral transmissibility of prion disorders across species as well as the intestinal pathophysiology of alimentary prion infections.24 Moreover, the blockage of such interactions through the use of anti-LRP/LR specific antibodies has been reported to reduce PrPSc endocytosis19 and thus these antibodies may serve as potential therapeutics to prevent infectious prion internalization and thereby prevent prion infections. It must be emphasized that the adhesion of prion proteins to cells is not solely dependent on the LRP/LR-PrPSc interactions;24 however, this interaction is of importance with regards to internalization and subsequent pathogenesis.We applied the aforementioned cell model to study the possible oral transmission of PrPBSE, PrPCWD and ovine PrPSc to cervids, cattle, swine and humans.24 The direct transmission of the aforementioned animal prion disorders to humans as a result of dietary exposure and the possible establishment of zoonotic diseases is of great public concern. It must however be emphasized that the study investigated the co-localization of LRP/LR and various prion strains and not the actual internalization process.PrPBSE was shown to co-localize with LRP/LR on human enterocytes24, thereby suggesting that PrPBSE is transmissible to humans via the oral route which is widely accepted as the manner by which variant CJD originated. This suspicion was previously investigated using a macaque model, which was successfully perorally infected by BSE-contaminated material and subsequently lead to the development of a prion disorder that resembles vCJD.25 These results, due to the evolutionary relatedness between macaques and humans, allowed researchers to confirm the oral transmissibility of PrPBSE to humans. PrPBSE may also potentially lead to prion disorder establishment in swine,24 livestock of great economic and social importance.The prion disorder affecting elk, mule deer and white-tailed deer is termed CWD. Cases of the disease are most prevalent in the US but are also evident in Canada and South Korea.26,27 As the infectious prion isoform is reported to be present in the blood28 and skeletal muscle,29 hunting, consumption of wild venison and contact with other animal products derived from CWD-infected elk and deer may thereby pose a public health risk. Our studies demonstrate that PrPCWD co-localizes with LRP/LR on human enterocytes24 thereby suggesting a possible oral transmissibilty of this TSE to humans. This is, however, inconsistent with results obtained during intra-cerebral inoculation of the brains and spinal cords of transgenic mice overexpressing the human cellular prion protein (PrPc),26,27 which is essential for TSE disease establishment and progression. Further, discrepancies have also been reported with respect to non-human primates, as squirrel monkeys have been successfully intracerebrally inoculated with mule-deer prion homogenates,30 while cynolmolgus macaques were resistant to infection.31 CWD has been transmitted to ferrets, minks and goats32 and as these animals may serve as domestic animals or livestock, secondary transmission from such animals to humans, through direct contact or ingestion of infected material, may be an additional risk factor that merits further scientific investigation.Ovine PrPSc co-localization with LRP/LR on human and bovine enterocytes may be indicative of the infectious agents'' ability to effect cross-species infections. The oral transmissibility of Scrapie has been confirmed in hamsters fed with sheep-scrapie-infected material.33The discrepancies with regards to the transmissibility of certain infectious prion proteins when assessed by different model systems may be due to the experimental transmission route employed. Oral exposure often results in significantly prolonged incubation times when compared to intracerebral inoculation techniques and thus failure of transgenic mice and normal experimental animals to develop disease phenotypes after being fed TSE-contaminated material may not necessarily indicate that the infection process failed.14 Apart from the route of infection, numerous other factors may influence transmission between species, including dose, PrP polymorphisms and genetic factors, the prion strain employed as well as the efficacy of prion transport to the CNS.34 The degree of homology between the PrPc protein in the animals serving as the infectious prion source and recipient has also been described as a feature limiting cross-species transmission.34 The negative results, as referred to above, obtained upon prion-protein inoculation of animal models may have resulted due to the slow rate at which the infectious prion induces conformational conversion of the endogenous PrPc in the animal cells and this in turn results in low levels of infectious prion replication and symptom development.27Furthermore, even in the event that certain prion disorders are not directly transmissible to humans, most are transmissible to at least a single species of domestic animal or livestock. The infectious agents properties may be altered in the secondary host such that it becomes transmissible to humans (reviewed in ref. 35). Thus, interspecies transmission between animals may indirectly influence human health.It is noteworthy to add that although the oral route of PrPSc transmission may result in prolonged incubation times, it may broaden the range of susceptible hosts. A common constituent of food is ferritin, a protein that is resistant to digestive enzyme hydrolysis and, due to its homology across species, it may serve as co-transporter of PrPSc and facilitate enterocyte internalization of the infectious prion.36 It may thus be proposed that prion internalization may occur via a ferritin-PrPSc complex even in the absence of co-localization between the infectious agent and LRP/LR such that many more cross-species infections (provided that the other infection factors are favorable) may be probable.37 In addition, digestive enzymes in the gastrointestinal tract facilitate PrPSc binding to the intestinal epithelium and subsequent intestinal uptake36 and thus depending on the individuals'' digestive processes, the susceptibility to infection and the rate of disease development may vary accordingly. As a result hereof, though laboratory experiments in cell-culture and animal models may render a particular prion disorder non-infectious to humans, this may not be true for all individuals.In lieu of the above statements, with particular reference to inconsistencies in reported results and the multiple factors influencing oral transmissibility of TSEs, further transmission studies are required to evaluate the zoonotic threat which CWD, BSE and Scrapie may pose through ingestion.  相似文献   

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

11.
Prion protein (PrP)-like molecule, doppel (Dpl), is neurotoxic in mice, causing Purkinje cell degeneration. In contrast, PrP antagonizes Dpl in trans, rescuing mice from Purkinje cell death. We have previously shown that PrP with deletion of the N-terminal residues 23–88 failed to neutralize Dpl in mice, indicating that the N-terminal region, particularly that including residues 23–88, may have trans-protective activity against Dpl. Interestingly, PrP with deletion elongated to residues 121 or 134 in the N-terminal region was shown to be similarly neurotoxic to Dpl, indicating that the PrP C-terminal region may have toxicity which is normally prevented by the N-terminal domain in cis. We recently investigated further roles for the N-terminal region of PrP in antagonistic interactions with Dpl by producing three different types of transgenic mice. These mice expressed PrP with deletion of residues 25–50 or 51–90, or a fusion protein of the N-terminal region of PrP with Dpl. Here, we discuss a possible model for the antagonistic interaction between PrP and Dpl.Key words: prion protein, doppel, neurotoxic signal, neurodegeneration, neuroprotection, prion diseaseThe normal prion protein, termed PrPC, is a membrane glycoprotein tethered to the outer cell surface via a glycosylphosphatidylinositol (GPI) anchor moiety.1,2 It is ubiquitously expressed in neuronal and non-neuronal tissues, with highest expression in the central nervous system, particularly in neurons.3 The physiological function of PrPC remains elusive. We and others have shown that PrPC functionally antagonizes doppel (Dpl), a PrP-like GPI-anchored protein with ∼23% identity in amino acid composition to PrP, protecting Dpl-induced neurotoxicity in mice.47 Dpl is encoded on Prnd located downstream of the PrP gene (Prnp) and expressed in the testis, heart, kidney and spleen of wild-type mice but not in the brain where PrPC is actively expressed.4,5,8 However, when ectopically expressed in brains, particularly in cerebellar Purkinje cells, Dpl exerts a neurotoxic activity, causing ataxia and Purkinje cell degeneration in Ngsk, Rcm0 and Zrch II lines of mice devoid of PrPC (Prnp0/0).4,9,10 In these mice, Dpl was abnormally controlled by the upstream Prnp promoter.4,5 This is due to targeted deletion of part of Prnp including a splicing acceptor of exon 3.11 Pre-mRNA starting from the residual exon1/2 of Prnp was abnormally elongated until the end of Prnd and then intergenically spliced between the residual Prnp exons 1/2 and the Prnd coding exons.4,5 As a result, Dpl was ectopically expressed under the control of the Prnp promoter in the brain, particularly in neurons including Purkinje cells.4,5 In contrast, in other Prnp0/0 lines, such as Zrch I and Npu, the splicing acceptor was intact, resulting in normal Purkinje cells without ectopic expression of Dpl in the brain.4The molecular mechanism of the antagonistic interaction between PrPC and Dpl remains unknown. We recently showed that the N-terminal half of PrPC includes elements that might mediate cis or trans protection against Dpl in mice, ameliorating Purkinje cell degeneration.12 We also showed that the octapeptide repeat (OR) region in the N-terminal domain is dispensable for PrPC to neutralize Dpl neurotoxicity in mice.12 Here, possible molecular mechanisms for the antagonism between PrPC and Dpl will be discussed.  相似文献   

12.
Yeast prions provide a powerful model system for examining prion formation and propagation in vivo. Yeast prion formation is driven primarily by amino acid composition, not by primary amino acid sequence. However, although yeast prion domains are consistently glutamine/asparagine-rich, they otherwise vary significantly in their compositions. Therefore, elucidating the exact compositional requirements for yeast prion formation has proven challenging. We have developed an in vivo method that allows for estimation of the prion propensity of each amino acid within the context of a yeast prion domain.1 Using these values, we are able to predict the prion-propensity of various glutamine/asparagine-rich yeast domains. These results provide insight into the basis for yeast prion formation, and may aid in the discovery of additional novel prion domains. Additionally, we examined whether amino acid composition could drive interactions between heterologous glutamine/asparagine-rich proteins.2 Although inefficient interactions between yeast prion domains have previously been observed, we found that one prion protein, Ure2, is able to interact with compositionally similar domains with unprecedented efficiency. This observation, combined with the growing number of yeast prions, suggests that a broad network of interactions between heterologous glutamine/asparagine-rich proteins may affect yeast prion formation.Key words: yeast, prion, amyloid, Ure2, Sup35The highly-studied yeast prion proteins Sup35, Ure2 and Rnq1, which form the [PSI+], [URE3] and [PIN+] prions, respectively, each contain a glutamine/asparagine (Q/N) rich domain that drives prion formation. 3 Prion formation is thought to involve conversion of the soluble proteins into an insoluble amyloid form.3 Four additional yeast prion proteins containing Q/N-rich prion-forming domains have recently been discovered, and domains with similar Q/N-content are over-represented in various eukaryotic genomes.48 This suggests that numerous other prion proteins may remain to be discovered, yet predicting which of these Q/N-rich domains is likely to form prions has proven difficult.Randomizing the order of the amino acids in either the Sup35 or Ure2 prion domains does not block prion formation, demonstrating that prion formation is driven primarily by amino acid composition, not primary sequence.9,10 We therefore explored two related questions. First, can amino acid composition be used to accurately predict prion propensity?1 Second, to what extent can compositionally similar Q/N-rich proteins interact during prion formation?2Surprisingly, although composition drives yeast prion formation, compositional similarity to known prion domains is a poor predictor of prion propensity. In addition to high Q/N content, yeast prion domains show an under-representation of both charged and hydrophobic residues (1,11 Alberti et al. recently identified the 100 yeast domains with greatest compositional similarity to the known prion domains, and tested each domain in four different assays for prion-like activity.4 This remarkable effort revealed numerous new potential prion domains; however, there was very little correlation between a domain’s compositional similarity to known prion domains and its prion-like activity.1,4 The most likely explanation for this unexpected finding is that because the yeast prion domains are likely not optimized for maximum prion propensity, some compositional deviations will increase prion propensity and some will decrease it. However, algorithms that predict prion propensity based on compositional similarity to known prion domains are predicated on the assumption that all compositional changes will reduce prion propensity. Instead, accurate prediction of prion propensity requires understanding how changes in amino acid composition affect prion propensity.

Table 1

Prion propensity, order propensity and prevalence for each amino acid
Amino acidPrion propensityaPrevalence in prion domainsbHydrophobicityc
Phenylalanine0.842.50.81
Isoleucine0.811.11.0
Valine0.811.50.97
Tyrosine0.787.80.36
Methionine0.671.70.71
Tryptophan0.6700.40
Cysteine0.4200.78
Serine0.1310.00.41
Asparagine0.08023.50.11
Glutamine0.06922.00.11
Glycine−0.03912.90.46
Leucine−0.0402.30.92
Threonine−0.122.10.42
Histidine−0.280.90.14
Alanine−0.403.90.70
Arganine−0.412.50.00
Glutamic Acid−0.611.50.11
Proline−1.201.70.32
Aspartic Acid−1.281.30.11
Lysine−1.580.70.07
Open in a separate windowaExperimentally determined prion propensity from ref. 1. Prion propensities were calculated as the natural log of the fold over/under-representation of the amino acid among prion-forming clones relative to the naïve library.bThe average percent prevalence among the Sup35, rnq1 and Ure2 prion domains.cFrom ref. 21, rescaled from 0 to 1. This scale is used by FoldIndex to calculate intrinsic folding propensity (IF) by the equation IF = 2.785 — <R> — 1.151, where <H> is the mean hydrophobicity and <R> is average net charge.20We therefore developed the first method to quantify the prion propensity of each amino acid within the context of a Q/N-rich prion domain (Fig. 1). We utilized Sup35-27, a scrambled version of Sup35 that forms prions with high efficiency. Using an oligonucleotide-based mutagenesis approach, we replaced eight consecutive codons within the SUP35-27 DNA sequence with (NNB)8, where N represents any of the four nucleotides and B represents any nucleotide except adenine. This generated a library of sequences in which all 20 amino acids should be present at each position within the mutated region. We then selected for the subset of proteins that maintained the ability to form prions. By comparing the naïve library to the prion-forming subset, we were able to determine the degree to which each amino acid was over- or under-represented among the prion-forming clones. These numbers were then used to generate a scale ranking the prion-propensity of all 20 amino acids (1,11,12 By contrast, hydrophobic residues were strongly overrepresented, and no detectable bias was seen for or against glutamines and asparagines.1 Similar results were seen at a second position, demonstrating that these biases were not an artifact of local interactions.Open in a separate windowFigure 1Mutagenesis method. Oligonucleotides were designed with a degenerate region, flanked by regions complementary to SUP35-27. These degenerate oligonucleotides were used to PCR amplify SUP35-27, generating a library of sequences in which eight codons were randomly mutated. Using plasmid shuffling, the library was used to replace wild-type Sup35 as the sole copy of Sup35 in the cell. Mutants were selected for prion formation by plating on medium lacking adenine. Isolates from the naïve library and the prion-forming subset were sequenced to determine which amino acids were over/under-represented among the prion-forming isolates.Many of these results can be rationalized based on the proposed structure of Sup35-27 amyloid fibrils. Amyloid fibrils are β-sheet-rich structures in which the β-strands are predominantly oriented perpendicular to the long axis of the fibril. Solid state NMR indicates that for Sup35-27, the β-strands adopt an inregister parallel alignment.13 This inregister alignment should strongly disfavor charged residues due to electrostatic repulsion. Proline residues are known β-sheet breakers, so it is not surprising that they would also be disfavored. Hydrophobic residues have been proposed to drive formation of other inregister parallel amyloid structures, so their high prion-propensities in our assay should not be a complete surprise.14Other aspects of our results are less easily explained based on proposed structures. X-ray diffraction studies of small fragments from Sup35 indicated that hydrogen bonding of in-register glutamines and asparagines stabilizes prion fibers, analogous to the polar zippers that have been proposed for poly-Q.15,16 However, we saw no bias in favor of glutamines and asparagines. This lack of bias in favor of glutamines and asparagines was particularly surprising based on the compositions of known yeast prion domains. Although high Q/N content is not an absolute requirement for a protein to act as a prion in yeast, Q/N residues are consistently over-represented in yeast prion domains (11 Likewise, although we observed a strong bias in favor of hydrophobic residues, and although tyrosines have been implicated in prion formation and propagation, hydrophobic residues in general are strongly under-represented among yeast prion domains (1,11,17,18This raises the question, why is there such a large discrepancy between the residues that most strongly promote prion formation and those that are actually present in yeast prion domains? The simplest explanation is that the composition of yeast prion domains may reflect functions of the domains other than prion formation. Bioinformatic analysis can reveal what compositional features are present in prion domain, but not why they are present. Nevertheless, it seems surprising that yeast prion domains could form prions when they are largely lacking in the most strongly prion-promoting residues. We propose that native state intrinsic disorder is essential to yeast prion formation, and can help explain the ability of yeast prion domains to form prions despite their relative lack of strongly prion-prone amino acids. The yeast prion domains are intrinsically disordered, and for Sup35, this structural flexibility seems to be important for prion formation.19 The yeast prion domains are biased towards residues that balance intrinsic disorder and prion-propensity. For most proteins, amyloid formation must compete with native-state structure; in the absence of this competition, yeast prion domains are able to form prions despite relatively modest prion propensities. Although hydrophobic residues have high prion propensity, large numbers of hydrophobic residues would reduce intrinsic disorder by promoting hydrophobic collapse, potentially creating a stable fold that would compete with amyloid formation.A significant question was whether mutagenesis results from short eightresidue segments could be used to predict prion propensities of entire prion domains. To test this, we scanned various proteins with a 41-amino-acid window size, calculating for each window prion propensity as the sum of the experimentally determined prion propensities for each amino acid in the window, and the order propensity using FoldIndex.20 FoldIndex is a simple algorithm for identifying disordered regions based on hydrophobicity (using the Kyte/Doolittle scale) and net charge.20,21 Strikingly, regions with similar predicted prion propensity to the yeast prion domains are common in non-prion proteins; however, these regions are consistently predicted to be ordered.1 By contrast, for Sup35 and Ure2, the prion domains are easily distinguishable as regions with positive prion propensity and negative FoldIndex order propensity (Fig. 2), supporting a role for intrinsic disorder in facilitating prion formation. Although the prion domains for the other yeast prion proteins are not as clearly defined, similar disordered and prion-prone regions are found in all of the yeast prion domains except that of Cyc8. By contrast, while Q/N-rich domains with low prion-like activity are consistently predicted to be disordered, they generally have low predicted prion propensity. By scoring proteins based on the predicted prion propensity of the most prion-prone region, we were able to distinguish with greater than 90% accuracy between Q/N-rich domains with and without prion-like activity. This is the first time that such prediction accuracy has been achieved for Q/N-rich domains.Open in a separate windowFigure 2Prion propensity maps. Ure2p (A) and Sup35 (B) were scanned using a window size of 41 amino acids, calculating for each window the average order propensity using FoldIndex and prion propensity as the sum of the experimentally determined prion propensities for each amino acid across the window. The prion domain (PFD) is shaded.These results provide insight into the differences between Q/N-rich and non-Q/N-rich amyloid proteins. Amyloid formation by non-Q/N-rich proteins is thought to be driven by short peptide stretches.22 Therefore, most amyloid prediction algorithms are designed to look for local regions of high prion propensity. However, our algorithm uses a relatively large 41-amino-acid window size. When we used a smaller window size, our algorithm lost the ability to distinguish between Q/N-rich domains with and without prion-like activity. Thus, the yeast prion domains are characterized by extended disordered regions of modest prion propensity, not local regions of high prion propensity. This explains why many amyloid prediction algorithms underestimate the amyloid propensity of the polar yeast prion domains.23An obvious question is whether mutagenesis data from a single protein can really be used to make such broad claims about the basis for prion formation by all Q/N-rich yeast prion proteins. Similarly, FoldIndex is a relatively simple disorder-prediction algorithm that may not be specifically optimized for Q/N-rich proteins. However, these concerns are somewhat allayed by the accurate predictions that our algorithm, when combined with FoldIndex, is able to make. The fact that mutagenesis data from a single Q/N-rich protein allows for accurate prediction of a broad range of other Q/N-rich proteins suggests that the basic mechanisms of prion formation are similar for the various Q/N-rich prion domains. There is little doubt that a broader data set derived from multiple yeast prion domains could improve prediction accuracy; likewise, other disorder-prediction algorithms might be better suited for analyzing yeast prions. Nevertheless, the prediction accuracy that we have achieved suggests that while such changes might modestly improve our prediction accuracy, they are unlikely to fundamentally change our general conclusions about yeast prion domains.Although our results provide an exciting first step towards understanding the sequence requirements for prion formation, numerous questions remain unanswered. For Sup35, separate regions of the prion domain are required for prion formation and prion propagation.24 Therefore, future experiments will be needed to dissect the distinct sequence requirements for prion formation versus propagation. Likewise, although numerous domains have been identified that can drive prion-like activity when inserted in the place of the Sup35 prion domain,4,25 flanking sequences outside of the Sup35 prion domain affect prion propagation.26 Therefore, there is no guarantee that such domains will be able to act as prions in their native contexts. Understanding the role of flanking sequences will be critical for predicting prion propensity of domains within their native contexts. Similarly, expression levels and patterns and cellular localization likely influence prion formation and propagation.The effects of primary sequence on prion propensity are also still unclear. The dominant effect of amino acid composition on prion formation has made it difficult to distinguish the more subtle effects of primary sequence. By delineating the compositional requirements for prion formation, our work should make it easier to identify primary sequence elements that affect prion propensity. Specifically, examination of outlier proteins not accurately predicted by our composition-based algorithm may reveal primary sequence features that promote or inhibit prion formation.One such aspect of primary sequence is already incorporated into our algorithm. In examining the 100 Q/N-rich proteins studied by Alberti et al. we discovered a subtle, but statistically significant bias in the distribution of proline residues.1,4 When prolines were present in Q/N-rich domains with prion-like activity, they tended to occur in clusters, while the prolines were more likely to be dispersed in domains without prion-like activity. This is not surprising, as prolines are known β-sheet breakers. If multiple prolines are present in single cluster, they will disrupt the β-sheet structure at a single location; by contrast, the same number of prolines dispersed throughout a sequence will result in multiple locations where the β-sheet structure is disrupted.Another potential example of the effects of primary sequence is seen in Cyc8, the only yeast prion protein incorrectly predicted by our algorithm not to form prions. Interestingly, the prion domain contains an imperfect (QA)32 repeat. Patterns of alternating polar and non-polar residues are thought to promote amyloid formation, although alanine has not generally been included among the non-polar residues when considering such patterns.27 Therefore, this primary sequence element may explain why our algorithm, which predominantly considers amino acid composition, predicts Cyc8 to have relatively low prion propensity.The high level of prediction accuracy achieved by our algorithm should also facilitate the identification of new prion proteins. The large number of yeast prion proteins, combined with the prevalence of Q/N-rich domains in eukaryotic genomes, suggests that similar prion-like structural conversions may be common in other organisms. However, only two prion proteins have been identified outside of Saccharomyces cerevisiae—the mammalian protein PrP and Podospora anserine protein HET-s—neither of which is Q/N-rich. Identification of new prion proteins has been hindered by the lack of accurate prediction algorithms. In yeast, genetic screens and compositional homology searches have led to the identification of novel prions, but with relatively low success rates.48 Understanding the compositional requirements for prion formation should improve this success rate, allowing for more targeted testing of potential prion proteins.As the list of Q/N-rich yeast prion proteins continues to grow, an obvious question will be, how do interactions between heterologous prion proteins affect prion formation and propagation? If amino acid composition is the predominant factor determining prion-like activity, can amino acid composition also drive interactions between heterologous prion proteins? Various interactions have been observed among the yeast prion proteins. Under normal cellular conditions, prion formation by Sup35 requires the presence of [PIN+].28 In vitro and in vivo evidence suggests that Rnq1 aggregates can act as imperfect templates for seeding Sup35 aggregation.29 Overexpression of various Q/N-rich proteins can substitute for [PIN+], allowing [PSI+] formation in cells lacking [PIN+]; this suggests that Sup35 can interact with a broad range of Q/N-rich proteins.30,31 [PIN+] also promotes [URE3] formation, while [PSI+] inhibits [URE3] formation.32,33 However, each of these interactions is quite inefficient; for example, Rnq1 aggregates are at least 50-fold less efficient than Sup35 aggregates at seeding Sup35 amyloid formation.29Therefore, we were surprised to discover that [URE3] prion formation can be stimulated with high efficiency by a variety of compositionally similar domains.2 Overexpression of the prion domain of Ure2 significantly increases [URE3] formation. Surprisingly, overexpression of each of the scrambled versions of Ure2 stimulates wild-type [URE3] formation with comparable efficiency.2 In vitro, amyloid aggregates formed by the scrambled Ure2s were able to efficiently seed amyloid aggregation by wild-type Ure2, suggesting that wild-type Ure2 is able to directly interact with scrambled Ure2 amyloid aggregates. There were limits to Ure2’s promiscuity, as overexpression of scrambled Sup35 prion domains did not promote [URE3] formation.To determine whether Ure2 could similarly interact with naturally occurring yeast proteins, we performed a genomic search to identify the five yeast protein fragments with greatest compositional similarity to the Ure2 prion domain.2 Four of these five fragments were able to stimulate wild-type [URE3] formation. The efficiency of this cross-seeding was unprecedented; one of the fragments, from Sap30, was able to stimulate [URE3] formation three-fold more efficiently than the wild-type Ure2 prion domain. These results raise the possibility that such interactions could be physiologically relevant, acting as regulators of either beneficial or deleterious amyloid formation. Interestingly, not all prion proteins appear to be as promiscuous as Ure2. Prion formation by Sup35 was not significantly affected by overexpression of scrambled versions of Sup35. Defining the sequence features that allow Ure2 to interact with compositionally similar domains, and defining the limits of Ure2’s promiscuity, will be critical for determining the extent to which similar interactions affect the normal physiology of yeast prions and mammalian amyloid diseases.Although it is clear that Ure2 is able to promiscuously interact with a broad range of compositionally similar proteins, our results say nothing about why such interactions occur. These interactions may have evolved to positively or negatively regulate prion formation. Alternatively, such prion-promoting interactions may simply be a byproduct of the natural function of the Ure2 prion domain. Although the exact function of the Ure2 prion domain is unknown, regions of intrinsic disorder domains are often used to recognize multiple binding partners.34 Promiscuous prion-promoting interactions may simply be an unintended consequence of such activity. Therefore, future studies will be needed to determine the function, if any, of Ure2’s promiscuity. Likewise, although our mutagenesis results clearly indicate that the compositions of the yeast prion domains reflect a balance of intrinsic disorder and prion propensity, our studies say nothing about why the yeast prion domains have evolved this balance. These compositional characteristics may have evolved for a reason unrelated to prion formation. Alternatively, the yeast prion domains may have specifically been optimized for the purpose of prion formation.  相似文献   

13.
Numerous prions (infectious proteins) have been identified in yeast that result from the conversion of soluble proteins into β-sheet-rich amyloid-like protein aggregates. Yeast prion formation is driven primarily by amino acid composition. However, yeast prion domains are generally lacking in the bulky hydrophobic residues most strongly associated with amyloid formation and are instead enriched in glutamines and asparagines. Glutamine/asparagine-rich domains are thought to be involved in both disease-related and beneficial amyloid formation. These domains are overrepresented in eukaryotic genomes, but predictive methods have not yet been developed to efficiently distinguish between prion and nonprion glutamine/asparagine-rich domains. We have developed a novel in vivo assay to quantitatively assess how composition affects prion formation. Using our results, we have defined the compositional features that promote prion formation, allowing us to accurately distinguish between glutamine/asparagine-rich domains that can form prion-like aggregates and those that cannot. Additionally, our results explain why traditional amyloid prediction algorithms fail to accurately predict amyloid formation by the glutamine/asparagine-rich yeast prion domains.Amyloid fibers are associated with a large number of neurodegenerative diseases and systemic amyloidoses. Amyloid fibrils are rich in a cross-beta quaternary structure in which β-strands are perpendicular to the long axis of the fibril (8).[URE3] and [PSI+] are the prion (infectious protein) forms of the Saccharomyces cerevisiae proteins Ure2 and Sup35, respectively (61). Formation of both prions involves conversion of the native proteins into an infectious, amyloid form. Ure2 and Sup35 have served as powerful model systems for examining the basis for amyloid formation and propagation. Both proteins possess a well-ordered functional domain responsible for the normal function of the protein, while a functionally and structurally separate glutamine/asparagine (Q/N)-rich intrinsically disordered domain is necessary and sufficient for prion aggregation and propagation (4, 26, 27, 52, 53). Both proteins can form multiple prion variants, which are distinguished by the efficiency of prion propagation and by the precise structure of the amyloid core (14, 54).Five other prion proteins have also been identified in yeast: Rnq1 (13, 46), Swi1 (15), Cyc8 (33), Mca1 (30), and Mot3 (1). Numerous other proteins, including New1, contain domains that show prion activity when inserted in place of the Sup35 prion-forming domain (PFD) (1, 42). Each of these prion proteins contains a Q/N-rich PFD. Similar Q/N-rich domains are overrepresented in eukaryotic genomes (28), raising the intriguing possibility that prion-like structural conversions by Q/N-rich domains may be common in other eukaryotes. However, we currently have little ability to predict whether a given Q/N-rich domain can form prions.A variety of algorithms have been developed to predict a peptide''s propensity to form amyloid fibrils based on its amino acid sequence, including BETASCAN (6), TANGO (17), Zyggregator (51), SALSA (62), and PASTA (55). These algorithms have been successful at identifying regions prone to amyloid aggregation and predicting the effects of mutations on aggregation propensity for many amyloid-forming proteins. However, they have generally been quite ineffective for Q/N-rich amyloid domains such as the yeast PFDs. For example, using the statistical mechanics-based algorithm TANGO (17), which predicts aggregation propensity based on a peptide''s physicochemical properties, Linding et al. found that the Sup35 and Ure2 PFDs both completely lack predicted β-aggregation nuclei (24). Similarly, yeast PFDs are generally lacking in the hydrophobic residues predicted by algorithms such as Zyggregator to nucleate amyloid formation.Why are these algorithms so effective for many amyloid-forming proteins but not for yeast PFDs? For most amyloid proteins, amyloid formation is driven by short hydrophobic protein stretches, and increased hydrophobicity is correlated with an increased amyloid aggregation propensity (34). In contrast, the yeast PFDs are all highly polar domains, due largely to the high concentration of Q/N residues and the lack of hydrophobic residues. High Q/N content is clearly not a requirement for a domain to act as a prion in yeast, since neither the mammalian prion protein PrP nor the Podospora anserina prion protein HET-s is Q/N rich, yet fragments from both proteins can act as prions in yeast (49, 50). However, the significant compositional differences between the yeast PFDs and most other amyloid/prion proteins suggest that there may be two distinct classes of amyloid-forming proteins driven by different types of interactions. Specifically, Q/N residues, which are predicted to have a relatively low amyloid propensity in the context of hydrophobic amyloid domains (34), may promote amyloid formation when present at sufficiently high density. Stacking of Q/N residues to form polar zippers has been proposed to stabilize amyloid fibrils (35). Consistent with this hypothesis, mutational studies of Sup35 indicate that Q/N residues are critical for driving [PSI+] formation (12), and expanded poly-Q or poly-N tracts are sufficient to drive amyloid aggregation (36, 63). Therefore, this paper examines the sequence features that allow the polar, Q/N-rich yeast PFDs to form prions.Mutational studies of the PFDs of Ure2 and Sup35 have shown that amino acid composition is the predominant feature driving prion formation (40, 41). Due to the unique compositional biases observed in the yeast PFDs, algorithms have been developed to identify potential PFDs based solely on amino acid composition (19, 28, 42). These algorithms are designed to produce a list of potential prion proteins that meet a specific set of criteria (such as high Q/N content) but are not able to predict the prion propensity of each member of the list or to predict the effects of mutations on prion formation. A recent study by Alberti et al. was the first to systematically test whether compositional similarity to known PFDs is sufficient to distinguish between Q/N-rich proteins that form prions and those that do not. They developed a hidden Markov model to identify domains that are compositionally similar to known PFDs and then analyzed the 100 highest-scoring Q/N-rich domains in a series of in vivo and in vitro assays (1). Remarkably, they discovered 18 proteins with prion-like activity in all assays. However, an equal number, including some of the domains with greatest compositional similarity to known PFDs, showed no prion-like activity.This inability to distinguish between Q/N-rich proteins that form prions and those that do not might seem to suggest that amino acid composition is not an accurate predictor of prion propensity. However, an alternative explanation is that known yeast PFDs are not an ideal training set for a composition-based prediction algorithm, since yeast prions are likely not optimized for maximal prion propensity. It is unclear whether yeast prion formation is a beneficial phenomenon providing a mechanism to regulate protein activity or a detrimental phenomenon analogous to human amyloid disease. [PSI+] can increase resistance to certain stress conditions (56), but the failure to observe [PSI+] in wild yeast strains (29) argues that beneficial [PSI+] formation is at most a rare event. If yeast prions are diseases, the PFDs certainly would not be optimized for maximum prion potential. If prion formation is a beneficial event allowing for rapid conversion between active and inactive states, the prion potential of the PFD would be optimized such that the frequencies of prion formation and loss would yield the optimal balance of prion and nonprion cells (25). Thus, specific residues might be excluded from yeast PFDs either because they inhibit prion formation or because they too strongly promote prion formation; bioinformatic analysis can reveal which residues are excluded from yeast PFDs but not why they are excluded. Accurate prediction of prion propensity requires understanding which deviations from known prion-forming compositions will promote prion formation and which will inhibit.We have therefore developed the first in vivo method to quantitatively determine the prion propensity for each amino acid in the context of a Q/N-rich PFD. As expected, we found proline and charged residues to be strongly inhibitory to prion formation; but surprisingly, despite being largely underrepresented in yeast PFDs, hydrophobic residues strongly promoted prion formation. Furthermore, although Q/N residues dominate yeast PFDs, prion propensity appears relatively insensitive to the exact number of Q/N residues. Using these data, we were able to distinguish with approximately 90% accuracy between Q/N-rich domains that can form prion-like aggregates and those that cannot. These experiments provide the first detailed insight into the compositional requirements for yeast prion formation and illuminate the different methods by which Q/N- and non-Q/N-rich amyloidogenic proteins aggregate.  相似文献   

14.
VERNALIZATION INSENSITIVE 3 (VIN3) encodes a PHD domain chromatin remodelling protein that is induced in response to cold and is required for the establishment of the vernalization response in Arabidopsis thaliana.1 Vernalization is the acquisition of the competence to flower after exposure to prolonged low temperatures, which in Arabidopsis is associated with the epigenetic repression of the floral repressor FLOWERING LOCUS C (FLC).2,3 During vernalization VIN3 binds to the chromatin of the FLC locus,1 and interacts with conserved components of Polycomb-group Repressive Complex 2 (PRC2).4,5 This complex catalyses the tri-methylation of histone H3 lysine 27 (H3K27me3),4,6,7 a repressive chromatin mark that increases at the FLC locus as a result of vernalization.4,710 In our recent paper11 we found that VIN3 is also induced by hypoxic conditions, and as is the case with low temperatures, induction occurs in a quantitative manner. Our experiments indicated that VIN3 is required for the survival of Arabidopsis seedlings exposed to low oxygen conditions. We suggested that the function of VIN3 during low oxygen conditions is likely to involve the mediation of chromatin modifications at certain loci that help the survival of Arabidopsis in response to prolonged hypoxia. Here we discuss the implications of our observations and hypotheses in terms of epigenetic mechanisms controlling gene regulation in response to hypoxia.Key words: arabidopsis, VIN3, FLC, hypoxia, vernalization, chromatin remodelling, survival  相似文献   

15.
Although propagation of Saccharomyces cerevisiae prions requires Hsp104 protein disaggregating activity, overproducing Hsp104 “cures” cells of [PSI+] prions. Earlier evidence suggests that the Hsp70 mutant Ssa1-21 impairs [PSI+] by a related mechanism. Here, we confirm this link by finding that deletion of STI1 both suppresses Ssa1-21 impairment of [PSI+] and blocks Hsp104 curing of [PSI+]. Hsp104''s tetratricopeptide repeat (TPR) interaction motif was dispensable for curing; however, cells expressing Sti1 defective in Hsp70 or Hsp90 interaction cured less efficiently, and the Hsp90 inhibitor radicicol abolished curing, implying that Sti1 acts in curing through Hsp70 and Hsp90 interactions. Accordingly, strains lacking constitutive or inducible Hsp90 isoforms cured at reduced rates. We confirm an earlier finding that elevating free ubiquitin levels enhances curing, but it did not overcome inhibition of curing caused by Hsp90 defects, suggesting that Hsp90 machinery is important for the contribution of ubiquitin to curing. We also find curing associated with cell division. Our findings point to crucial roles of Hsp70, Sti1, and Hsp90 for efficient curing by overexpressed Hsp104 and provide evidence supporting the earlier suggestion that destruction of prions by protein disaggregation does not adequately explain the curing.Saccharomyces cerevisiae prions are self-replicating misfolded forms of normal cellular proteins. They are believed to propagate as amyloid, which is a highly ordered fibrous aggregate. What triggers prion formation is uncertain, but in order to be maintained in an expanding yeast population, prions must grow, replicate, and be transmitted to daughter cells during cell division. Growth occurs when soluble protein joins the fiber ends and is converted into the prion form (30, 52, 58). Replication is associated with fragmentation of prion polymers, which generates new prions from preexisting material (37, 50). Transmission is believed to occur by passive diffusion of prions with cytoplasm (57).Although it is uncertain to what extent cellular factors influence growth or transmission of prions, it is clear that the Hsp104 disaggregation machinery is necessary for prion replication (10, 17, 55, 70). Hsp104 is a hexameric AAA+ chaperone that protects cells from a variety of stresses by resolubilizing proteins from aggregates (24, 25, 53). With help from Hsp70 and Hsp40, it extracts monomers from aggregates and extrudes them through its central pore (24, 41, 68). This machinery could act in prion replication by extracting monomers from amyloid fibers (29, 68), which would destabilize the fibers, causing them to break into more numerous pieces that each can continue to propagate the prion.Paradoxically, overexpressing Hsp104 very efficiently “cures” cells of the [PSI+] prion, which is composed of the translation termination factor Sup35 (10). A widely held view of this curing is that elevating the cellular protein disaggregation activity causes complete destruction of prions. However, elevating Hsp104 has little or no effect on most other amyloidogenic prions (15, 16, 38, 47, 54, 66), although it can be inferred to cure [MCA] prions in cells also propagating a prion of an Mca1-Sup35 fusion (49). Together, these results suggest that prions of Sup35, and perhaps those of Mca1, are particularly sensitive to Hsp104 disaggregation activity. Alternatively, something in addition to or other than a simple increase in protein disaggregation is involved in the curing.Although protein disaggregation activity of Hsp104 is required for both thermotolerance and prion propagation, we and others have identified mutations in Hsp104 that affect these processes separately (27, 32, 39, 60). The ability of Hsp104 to thread proteins through its central pore, however, is required for both processes (29, 41, 68), so this distinction in Hsp104 function could be due to differences in how Hsp104 interacts with amorphous aggregates of thermally denatured proteins and highly ordered prion aggregates or with cofactors that interact with the different prions as substrates. In any scenario, efficiency and specificity of Hsp104 function are affected by interactions with other components of the disaggregation machinery, in particular the Hsp70s and Hsp40s, which are believed to interact first with substrates to facilitate action of Hsp100 family disaggregases (2, 71, 72).Increasing expression of either ubiquitin (Ub) or Ssb, an Hsp70 that has roles in protein translation and proteasome degradation, enhances Hsp104 curing of [PSI+] (3, 11, 12). Predictably, reducing expression of either of them reduces curing efficiency. The mechanisms underlying these effects are unknown, but the combined effects of Ssb and Ub are additive, suggesting that they act in different pathways. The role of Ub is indirect, as Sup35 is neither ubiquitylated nor degraded during curing. Whether other chaperones are involved in the effects of Ub on curing has not been investigated.Earlier we isolated a mutant of the Hsp70 Ssa1, designated Ssa1-21, that weakens and destabilizes [PSI+] propagation (33). We later isolated several Hsp104 mutants that suppress this antiprion effect (29). The Hsp104 mutants retain normal functions in thermotolerance, protein disaggregation, and prion propagation, but when overexpressed, they are unable to cure [PSI+], even in wild-type cells. These findings argue against a specific hypersensitivity of [PSI+] to disaggregation and support the notion that something distinct from or in addition to complete destruction of prions is involved in the curing. They also imply that Ssa1-21 and elevated Hsp104 inhibit [PSI+] prions by similar mechanisms. A prediction from this conclusion is that other suppressors of Ssa1-21 will also inhibit curing of [PSI+] by overexpressed Hsp104. Indeed, we find here that alterations that suppress Ssa1-21 inhibition of [PSI+] do interfere with curing of [PSI+] by overexpressed Hsp104. We also provide evidence that Hsp90 has a critical role in this curing and that the ability of Ub to enhance curing depends on proper function of Hsp90 machinery.  相似文献   

16.
The pathogenicity of Clostridium difficile (C. difficile) is mediated by the release of two toxins, A and B. Both toxins contain large clusters of repeats known as cell wall binding (CWB) domains responsible for binding epithelial cell surfaces. Several murine monoclonal antibodies were generated against the CWB domain of toxin A and screened for their ability to neutralize the toxin individually and in combination. Three antibodies capable of neutralizing toxin A all recognized multiple sites on toxin A, suggesting that the extent of surface coverage may contribute to neutralization. Combination of two noncompeting antibodies, denoted 3358 and 3359, enhanced toxin A neutralization over saturating levels of single antibodies. Antibody 3358 increased the level of detectable CWB domain on the surface of cells, while 3359 inhibited CWB domain cell surface association. These results suggest that antibody combinations that cover a broader epitope space on the CWB repeat domains of toxin A (and potentially toxin B) and utilize multiple mechanisms to reduce toxin internalization may provide enhanced protection against C. difficile-associated diarrhea.Key words: Clostridium difficile, toxin neutralization, therapeutic antibody, cell wall binding domains, repeat proteins, CROPs, mAb combinationThe most common cause of nosocomial antibiotic-associated diarrhea is the gram-positive, spore-forming anaerobic bacillus Clostridium difficile (C. difficile). Infection can be asymptomatic or lead to acute diarrhea, colitis, and in severe instances, pseudomembranous colitis and toxic megacolon.1,2The pathological effects of C. difficile have long been linked to two secreted toxins, A and B.3,4 Some strains, particularly the virulent and antibiotic-resistant strain 027 with toxinotype III, also produce a binary toxin whose significance in the pathogenicity and severity of disease is still unclear.5 Early studies including in vitro cell-killing assays and ex vivo models indicated that toxin A is more toxigenic than toxin B; however, recent gene manipulation studies and the emergence of virulent C. difficile strains that do not express significant levels of toxin A (termed “A B+”) suggest a critical role for toxin B in pathogenicity.6,7Toxins A and B are large multidomain proteins with high homology to one another. The N-terminal region of both toxins enzymatically glucosylates small GTP binding proteins including Rho, Rac and CDC42,8,9 leading to altered actin expression and the disruption of cytoskeletal integrity.9,10 The C-terminal region of both toxins is composed of 20 to 30 residue repeats known as the clostridial repetitive oligopeptides (CROPs) or cell wall binding (CWB) domains due to their homology to the repeats of Streptococcus pneumoniae LytA,1114 and is responsible for cell surface recognition and endocytosis.12,1517C. difficile-associated diarrhea is often, but not always, induced by antibiotic clearance of the normal intestinal flora followed by mucosal C. difficile colonization resulting from preexisting antibiotic resistant C. difficile or concomitant exposure to C. difficile spores, particularly in hospitals. Treatments for C. difficile include administration of metronidazole or vancomycin.2,18 These agents are effective; however, approximately 20% of patients relapse. Resistance of C. difficile to these antibiotics is also an emerging issue19,20 and various non-antibiotic treatments are under investigation.2025Because hospital patients who contract C. difficile and remain asymptomatic have generally mounted strong antibody responses to the toxins,26,27 active or passive immunization approaches are considered hopeful avenues of treatment for the disease. Toxins A and B have been the primary targets for immunization approaches.20,2833 Polyclonal antibodies against toxins A and B, particularly those that recognize the CWB domains, have been shown to effectively neutralize the toxins and inhibit morbidity in rodent infection models.31 Monoclonal antibodies (mAbs) against the CWB domains of the toxins have also demonstrated neutralizing capabilities; however, their activity in cell-based assays is significantly weaker than that observed for polyclonal antibody mixtures.3336We investigated the possibility of creating a cocktail of two or more neutralizing mAbs that target the CWB domain of toxin A with the goal of synthetically re-creating the superior neutralization properties of polyclonal antibody mixtures. Using the entire CWB domain of toxin A, antibodies were raised in rodents and screened for their ability to neutralize toxin A in a cell-based assay. Two mAbs, 3358 and 3359, that (1) both independently demonstrated marginal neutralization behavior and (2) did not cross-block one another from binding toxin A were identified. We report here that 3358 and 3359 use differing mechanisms to modify CWB-domain association with CHO cell surfaces and combine favorably to reduce toxin A-mediated cell lysis.  相似文献   

17.
A role for SR proteins in plant stress responses   总被引:1,自引:0,他引:1  
  相似文献   

18.
Here we announce the complete genome sequence of Croceibacter atlanticus HTCC2559T, which was isolated by high-throughput dilution-to-extinction culturing from the Bermuda Atlantic Time Series station in the Western Sargasso Sea. Strain HTCC2559T contained genes for carotenoid biosynthesis, flavonoid biosynthesis, and several macromolecule-degrading enzymes. The genome confirmed physiological observations of cultivated Croceibacter atlanticus strain HTCC2559T, which identified it as an obligate chemoheterotroph.The phylum Bacteroidetes comprises 6 to ∼30% of total bacterial communities in the ocean by fluorescence in situ hybridization (8-10). Most marine Bacteroidetes are in the family Flavobacteriaceae, most of which are aerobic respiratory heterotrophs that form a well-defined clade by 16S rRNA phylogenetic analyses (4). The members of this family are well known for degrading macromolecules, including chitin, DNA, cellulose, starch, and pectin (17), suggesting their environmental roles as detritus decomposers in the ocean (6). Marine Polaribacter and Dokdonia species in the Flavobacteriaceae have also shown to have photoheterotrophic metabolism mediated by proteorhodopsins (11, 12).Several strains of the family Flavobacteriaceae were isolated from the Sargasso Sea and Oregon coast, using high-throughput culturing approaches (7). Croceibacter atlanticus HTCC2559T was cultivated from seawater collected at a depth of 250 m from the Sargasso Sea and was identified as a new genus in the family Flavobacteriaceae based on its 16S rRNA gene sequence similarities (6). Strain HTCC2559T met the minimal standards for genera of the family Flavobacteriaceae (3) on the basis of phenotypic characteristics (6).Here we report the complete genome sequence of Croceibacter atlanticus HTCC2559T. The genome sequencing was initiated by the J. Craig Venter Institute as a part of the Moore Foundation Microbial Genome Sequencing Project and completed in the current announcement. Gaps among contigs were closed by Genotech Co., Ltd. (Daejeon, Korea), using direct sequencing of combinatorial PCR products (16). The HTCC2559T genome was analyzed with a genome annotation system based on GenDB (14) at Oregon State University and with the NCBI Prokaryotic Genomes Automatic Annotation Pipeline (15, 16).The HTCC2559T genome is 2,952,962 bp long, with 33.9 mol% G+C content, and there was no evidence of plasmids. The number of protein-coding genes was 2,715; there were two copies of the 16S-23S-5S rRNA operon and 36 tRNA genes. The HTCC2559T genome contained genes for a complete tricarboxylic acid cycle, glycolysis, and a pentose phosphate pathway. The genome also contained sets of genes for metabolic enzymes involved in carotenoid biosynthesis and also a serine/glycine hydroxymethyltransferase, which is often associated with the assimilatory serine cycle (13). The potential for HTCC2559T to use bacterial type III polyketide synthase (PKS) needs to be confirmed because this organism had a naringenin-chalcone synthase (CHS) or chalcone synthase (EC 2.3.1.74), a key enzyme in flavonoid biosynthesis. CHS initiates the addition of three molecules of malonyl coenzyme A (malonyl-CoA) to a starter CoA ester (e.g., 4-coumaroyl-CoA) (1) and takes part in a few bacterial type III polyketide synthase systems (1, 2, 5, 18).The complete genome sequence confirmed that strain HTCC2559T is an obligate chemoheterotroph because no genes for phototrophy were found. As expected from physiological characteristics (6), the HTCC2559T genome contained a set of genes coding for enzymes required to degrade high-molecular-weight compounds, including peptidases, metallo-/serine proteases, pectinase, alginate lyases, and α-amylase.  相似文献   

19.
In young Arabidopsis seedlings, retrograde signaling from plastids regulates the expression of photosynthesis-associated nuclear genes in response to the developmental and functional state of the chloroplasts. The chloroplast-located PPR protein GUN1 is required for signalling following disruption of plastid protein synthesis early in seedling development before full photosynthetic competence has been achieved. Recently we showed that sucrose repression and the correct temporal expression of LHCB1, encoding a light-harvesting chlorophyll protein associated with photosystem II, are perturbed in gun1 mutant seedlings.1 Additionally, we demonstrated that in gun1 seedlings anthocyanin accumulation and the expression of the “early” anthocyanin-biosynthesis genes is perturbed. Early seedling development, predominantly at the stage of hypocotyl elongation and cotyledon expansion, is also affected in gun1 seedlings in response to sucrose, ABA and disruption of plastid protein synthesis by lincomycin. These findings indicate a central role for GUN1 in plastid, sucrose and ABA signalling in early seedling development.Key words: ABA, ABI4, anthocyanin, chloroplast, GUN1, retrograde signalling, sucroseArabidopsis seedlings develop in response to light and other environmental cues. In young seedlings, development is fuelled by mobilization of lipid reserves until chloroplast biogenesis is complete and the seedlings can make the transition to phototrophic growth. The majority of proteins with functions related to photosynthesis are encoded by the nuclear genome, and their expression is coordinated with the expression of genes in the chloroplast genome. In developing seedlings, retrograde signaling from chloroplasts to the nucleus regulates the expression of these nuclear genes and is dependent on the developmental and functional status of the chloroplast. Two classes of gun (genomes uncoupled) mutants defective in retrograde signalling have been identified in Arabidopsis: the first, which comprises gun2–gun5, involves mutations in genes encoding components of tetrapyrrole biosynthesis.2,3 The other comprises gun1, which has mutations in a nuclear gene encoding a plastid-located pentatricopeptide repeat (PPR) protein with an SMR (small MutS-related) domain near the C-terminus.4,5 PPR proteins are known to have roles in RNA processing6 and the SMR domain of GUN1 has been shown to bind DNA,4 but the specific functions of these domains in GUN1 are not yet established. However, GUN1 has been shown to be involved in plastid gene expression-dependent,7 redox,4 ABA1,4 and sucrose signaling,1,4,8 as well as light quality and intensity sensing pathways.911 In addition, GUN1 has been shown to influence anthocyanin biosynthesis, hypocotyl extension and cotyledon expansion.1,11  相似文献   

20.
Plant defensins are small, highly stable, cysteine-rich peptides that constitute a part of the innate immune system primarily directed against fungal pathogens. Biological activities reported for plant defensins include antifungal activity, antibacterial activity, proteinase inhibitory activity and insect amylase inhibitory activity. Plant defensins have been shown to inhibit infectious diseases of humans and to induce apoptosis in a human pathogen. Transgenic plants overexpressing defensins are strongly resistant to fungal pathogens. Based on recent studies, some plant defensins are not merely toxic to microbes but also have roles in regulating plant growth and development.Key words: defensin, antifungal, antimicrobial peptide, development, innate immunityDefensins are diverse members of a large family of cationic host defence peptides (HDP), widely distributed throughout the plant and animal kingdoms.13 Defensins and defensin-like peptides are functionally diverse, disrupting microbial membranes and acting as ligands for cellular recognition and signaling.4 In the early 1990s, the first members of the family of plant defensins were isolated from wheat and barley grains.5,6 Those proteins were originally called γ-thionins because their size (∼5 kDa, 45 to 54 amino acids) and cysteine content (typically 4, 6 or 8 cysteine residues) were found to be similar to the thionins.7 Subsequent “γ-thionins” homologous proteins were indentified and cDNAs were cloned from various monocot or dicot seeds.8 Terras and his colleagues9 isolated two antifungal peptides, Rs-AFP1 and Rs-AFP2, noticed that the plant peptides'' structural and functional properties resemble those of insect and mammalian defensins, and therefore termed the family of peptides “plant defensins” in 1995. Sequences of more than 80 different plant defensin genes from different plant species were analyzed.10 A query of the UniProt database (www.uniprot.org/) currently reveals publications of 371 plant defensins available for review. The Arabidopsis genome alone contains more than 300 defensin-like (DEFL) peptides, 78% of which have a cysteine-stabilized α-helix β-sheet (CSαβ) motif common to plant and invertebrate defensins.11 In addition, over 1,000 DEFL genes have been identified from plant EST projects.12Unlike the insect and mammalian defensins, which are mainly active against bacteria,2,3,10,13 plant defensins, with a few exceptions, do not have antibacterial activity.14 Most plant defensins are involved in defense against a broad range of fungi.2,3,10,15 They are not only active against phytopathogenic fungi (such as Fusarium culmorum and Botrytis cinerea), but also against baker''s yeast and human pathogenic fungi (such as Candida albicans).2 Plant defensins have also been shown to inhibit the growth of roots and root hairs in Arabidopsis thaliana16 and alter growth of various tomato organs which can assume multiple functions related to defense and development.4  相似文献   

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