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 78 毫秒
1.
2.
The study of fungal prion proteins affords remarkable opportunities to elucidate both intragenic and extragenic effectors of prion propagation. The yeast prion protein Sup35 and the self-perpetuating [PSI+] prion state is one of the best characterized fungal prions. While there is little sequence homology among known prion proteins, one region of striking similarity exists between Sup35p and the mammalian prion protein PrP. This region is comprised of roughly five octapeptide repeats of similar composition. The expansion of the repeat region in PrP is associated with inherited prion diseases. In order to learn more about the effects of PrP repeat expansions on the structural properties of a protein that undergoes a similar transition to a self-perpetuating aggregate, we generated chimeric Sup35-PrP proteins. Using both in vivo and in vitro systems we described the effect of repeat length on protein misfolding, aggregation, amyloid formation and amyloid stability. We found that repeat expansions in the chimeric prion proteins increase the propensity to initiate prion propagation and enhance the formation of amyloid fibers without significantly altering fiber stability.Key words: prion, yeast, sup35, PrP, nonsense suppression, translation termination, amyloid, repeatWe recently described a novel chimeric prion system that was designed to elucidate the consequences of one class of inherited prion disease mutations on protein folding.1,2 We created a fusion between the mammalian prion protein PrP and the yeast prion protein Sup35p (Fig. 1). Sup35p is an essential translation termination factor in yeast. Interestingly, the majority of the protein can be sequestered into a self-propagating aggregate, the [PSI+] prion.3 Remarkably, when yeast are grown in normal laboratory conditions, the [PSI+] prion is not detrimental. In fact, the biological consequences of the switch from the [psi−] non-prion state to the [PSI+] prion state may be beneficial in terms of adaptation and evolution.4 Importantly, the prion state of Sup35p can be readily detected in vivo by monitoring the reduced function of the translation termination factor when the protein is propagating as a prion aggregate.3 In addition, several methods have been developed to not only follow the propagation of the prion, but also to control the propagation and promote prion induction and loss (curing).5 Therefore, in addition to simply being a fascinating biological problem in of itself, the [PSI+] prion in yeast affords the ability to further elucidate both intragenic and extragenic effectors of prion biology.Open in a separate windowFigure 1Schematic representation of the yeast protein Sup35p and the mammalian prion protein PrP highlighting the position of the oligopeptide repeat domain (ORD). The amino acid sequence represents the consensus for a single repeat. Numbers shown represent the amino acid position of the beginning and the end of each ORD. The numbers above the schematic represent the original PrP amino acid positioning and the numbers below represent the original Sup35p amino acid sequence positions.Several prions have now been identified and interestingly, there is little sequence homology between the proteins to suggest that only one type of sequence can form a self-propagating aggregate.68 In vitro studies suggest that many proteins can form amyloids under the appropriate conditions.9 The fact that only a small percentage of proteins propagate as prions in vivo may be partly a consequence of physiological conditions being adequate to promote amyloid formation with those particular sequences. It is unclear what the precise distinction between prion and amyloid is at this time, but localization alone may preclude some amyloidogenic proteins from being “prion proteins” per se.10The sequence context that permits a protein to adopt a prion conformation in vivo is unclear. Several of the identified prion proteins have a domain that is enriched in glutamine and asparagine (Q/N) residues, but this is not true of all prion proteins.7 Our recent study demonstrates that the Q/N character of the Sup35p prion-forming domain can be significantly reduced, yet still propagate as a prion.1 This was also found recently in another prion protein chimera created and expressed in yeast.6 These studies suggest that the lack of stable secondary structure may be one of the defining features of a prion-forming domain. One of the striking sequence similarities that does exist between two prion proteins occurs in an oligopeptide repeat region found in Sup35p and PrP.11 Previous data clearly demonstrated that the Sup35p repeats are important for [PSI+] prion propagation.1215 The deletion of a single repeat from the wild type SUP35 sequence results in the loss of normal [PSI+] prion propagation.12 Moreover, the addition of two extra repeats of Sup35p sequence served to enhance the formation of the [PSI+] prion.13 The expansion of the analogous repeat domain in the mammalian prion protein PrP is associated with an inherited form of prion disease.16 Since the repeat regions of Sup35p and PrP are similar in size and character, we wanted to determine if the Sup35p oligopeptide repeat region could be substituted with that of PrP. Indeed, the PrP repeats in the context of Sup35p supported the propagation of the [PSI+] prion in yeast.1,17 Strikingly, we found phenotypic changes that occurred in a repeat length-dependent manner that suggested that the repeat expansions associated with disease result in an increase in the aggregation propensity but do not necessarily dictate only one type of aggregate structure.1More recently, we verified some of these results in vitro.2 These data are in agreement with other studies on the effect of repeat expansions.18,19 Taking the analysis one step further, we demonstrated that the stability of the amyloid fibers formed with the repeat-expanded proteins did not differ significantly. A very interesting observation that we made was that the formation of amyloid fibers by the longest repeat-expanded chimera (SP14NM) followed drastically different kinetics compared to the chimera containing the wild type number of repeats (SP5NM).2 In unseeded reactions, SP14NM did not show a lag phase during the course of fiber formation whereas SP5NM displayed a characteristic lag phase. Furthermore, the morphology of the amyloid fibers visualized by EM was different between SP14NM and SP5NM. SP14NM fibers were curvy and clumped but SP5NM fibers were long and straight. The correlation between the kinetics and the morphology of amyloid formation of SP14NM and SP5NM is reminiscent of fibers formed by β2-microglobulin (β2m) protein in different conditions.20 At pH 3.6, β2m formed curvy, worm-like fibers with no apparent lag phase. In contrast, long, straight fibers were formed at pH 2.5 and had a distinct lag phase. Analysis of the β2m fibers formed at pH 3.6 using mass spectrometric techniques identified species ranging from monomer to 13-mer. This suggested that the fibers were formed by monomer addition. On the other hand, oligomers larger than tetramers were not formed during fiber formation at pH 2.5. Based on these data the authors propose that β2m forms fibers in a nucleation-independent manner at pH 3.6, but fiber formation at pH 2.5 follows a nucleation-dependent mechanism. We suggest that the mechanism underlying SP5NM and repeat-expanded SP14NM fiber formation is similar to β2m fibers formed at pH 2.5 and pH 3.6, respectively. It will be interesting to determine if disease-associated mutations in amyloidogenic proteins alter the pathway whereby amyloid formation occurs and how that process plays a role in pathogenesis.In our in vivo study,1 we highlighted a unique feature of the longest Sup35-PrP chimera that related to the ability of the protein to adopt multiple self-perpetuating prion conformations more readily than wild type Sup35p. We suggest that this may be an important aspect of prion biology as it relates to inherited disease. If the repeat-expanded proteins can adopt multiple conformations that aggregate, then that may contribute to the large amount of variation observed in pathology and disease progression in this class of inherited prion diseases.21,22We also found that the spontaneous conversion of the repeat-expanded Sup35-PrP chimera into a prion state was significantly increased. However, this conversion required another aggregated protein in vivo, the [RNQ+] prion. In vitro, the prion-forming domain of the chimera showed a similar trend with the longer repeat lengths enhancing the ability of the protein to form amyloid fibers. The chimera with repeat expansions (8, 11 or 14 repeats) formed fibers very quickly as compared to that with the wild type number of repeats (5). While this correlates with the in vivo data in that both systems demonstrate an increased level of conversion with the repeat expansion, the systems are very different with respect to their requirement for a different “seed” to initiate the prion conversion. So, how does the [RNQ+] prion influence [PSI+]? At the moment, that isn''t entirely clear. Susan Liebman and colleagues discovered another epigenetic factor in yeast, [PIN+], which was important for the de novo induction of [PSI+].2325 Several years later, the [RNQ+] prion26 was found to be that factor in the commonly used [PSI+] laboratory strains, but they also found that the overexpression of other proteins could reproduce the effect.25 Hence, [RNQ+] can be [PIN+], and may be the primary epigenetic element that influences [PSI+] induction in yeast, but need not be in every case. Two models were proposed to explain the ability of [RNQ+] to influence the induction of [PSI+].25,27 One suggested that there is a direct templating effect where the aggregated state of the Rnq1 protein in the [RNQ+] prion serves as a seed for the direct physical association and aggregation of Sup35p and initiates [PSI+]. The second postulated that there is an inhibitor of aggregation in cells that is titrated out by the presence of another aggregated protein. Recent experimental evidence suggests that the templating model may explain at least part of the mechanism of action behind the [RNQ+] prion inducing the formation of [PSI+].28,29Why is [RNQ+] required for the in vivo conversion of the repeatexpanded chimera that forms amyloid on its own very efficiently in vitro? Interestingly, we found that the [RNQ+] prion per se is not required. We overexpressed the Rnq1 protein from a constitutive high promoter (pGPD-RNQ1) and found that Rnq1p aggregated in the cells but did not induce the [RNQ+] prion. That is, the cells were still [rnq−] and did not genetically transmit the aggregated state of the protein. However, even these non-prion aggregates of Rnq1p served to enhance the induction of the chimeric prions. Therefore, either the [RNQ+] prion or an aggregate of Rnq1 protein is sufficient, which is in line with previous studies that demonstrated that some proteins that aggregate when overexpressed can also enhance the induction of [PSI+].25 Also of note, recent data suggests that the requirement of [RNQ+] for the induction of Sup35p aggregation in vivo can be overcome by very long polyglutamine or glutamine/tyrosine stretches fused to the non-prion forming domain of Sup35p.30 These fusions may alter protein-protein interactions or destabilize the non-prion structure of Sup35p in such a manner that the [RNQ+] prion seed is no longer required to form [PSI+] de novo. Indeed, the non-polymerizing state of some of the fusion proteins was shown to be very unstable.So, what is the important difference between our in vitro and in vivo systems in the prion conversion? Obviously there are many candidates. First, the full length Sup35 protein may alter the conversion properties since a large part of the molecule is the structured C terminal domain. The C terminal domain may influence the initiation of prion propagation in vivo and that is not a factor in the in vitro system. Second, the influences of co-translational folding and potentially some initial unfolding of the prion-forming domain are not present since the in vitro system starts with denatured protein. Third, the environmental influences are clearly different. The molecular crowding effects and chaperones that are required for prion propagation in vivo are not required for the formation of amyloid in vitro. Finally, it is unclear if amyloid structures similar to those formed with the prion-forming domain in vitro actually exist in yeast. Certainly there is some correlation between the structures since aggregated Sup35 protein from [PSI+] cell lysates can seed amyloid formation in vitro31,32 and the fibers formed in vitro can be transformed into [psi−] cells and cause conversion to [PSI+].33 Nevertheless, we find it interesting that the expansion of the repeat region can have a tremendous effect on amyloid formation in vitro yet still cannot overcome the requirement for [RNQ+] for conversion in vivo. The presence of co-aggregating or cross-seeding proteins may play a role in the sporadic appearance or progression of neurodegenerative diseases and the interconnected yeast prions [RNQ+] and [PSI+] may provide a model system for elucidating the mechanism underlying such effects.  相似文献   

3.
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   

4.
5.
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   

6.
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   

7.
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   

8.
9.
10.
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   

11.
Amyloid fibrils share a structural motif consisting of highly ordered β-sheets aligned perpendicular to the fibril axis.1, 2 At each fibril end, β-sheets provide a template for recruiting and converting monomers.3 Different amyloid fibrils often co-occur in the same individual, yet whether a protein aggregate aids or inhibits the assembly of a heterologous protein is unclear. In prion disease, diverse prion aggregate structures, known as strains, are thought to be the basis of disparate disease phenotypes in the same species expressing identical prion protein sequences.47 Here we explore the interactions reported to occur when two distinct prion strains occur together in the central nervous system.Key words: prion, prions, strain, TSE, interaction, amyloid, LCP, neurodegeneration, aggregation  相似文献   

12.
13.
In our recent paper in the Plant Journal, we reported that Arabidopsis thaliana lysophospholipase 2 (lysoPL2) binds acyl-CoA-binding protein 2 (ACBP2) to mediate cadmium [Cd(II)] tolerance in transgenic Arabidopsis. ACBP2 contains ankyrin repeats that have been previously shown to mediate protein-protein interactions with an ethylene-responsive element binding protein (AtEBP) and a farnesylated protein 6 (AtFP6). Transgenic Arabidopsis ACBP2-overexpressors, lysoPL2-overexpressors and AtFP6-overexpressors all display enhanced Cd(II) tolerance, in comparison to wild type, suggesting that ACBP2 and its protein partners work together to mediate Cd(II) tolerance. Given that recombinant ACBP2 and AtFP6 can independently bind Cd(II) in vitro, they may be able to participate in Cd(II) translocation. The binding of recombinant ACBP2 to [14C]linoleoyl-CoA and [14C]linolenoyl-CoA implies its role in phospholipid repair. In conclusion, ACBP2 can mediate tolerance to Cd(II)-induced oxidative stress by interacting with two protein partners, AtFP6 and lysoPL2. Observations that ACBP2 also binds lysophosphatidylcholine (lysoPC) in vitro and that recombinant lysoPL2 degrades lysoPC, further confirm an interactive role for ACBP2 and lysoPL2 in overcoming Cd(II)-induced stress.Key words: acyl-CoA-binding protein, cadmium, hydrogen peroxide, lysophospholipase, oxidative stressAcyl-CoA-binding proteins (ACBP1 to ACBP6) are encoded by a multigene family in Arabidopsis thaliana.1 These ACBP proteins are well studied in Arabidopsis in comparison to other organisms,14 and are located in various subcellular compartments.1 Plasma membranelocalized ACBP1 and ACBP2 contain ankyrin repeats that have been shown to function in protein-protein interactions.5,6 ACBP1 and ACBP2 which share 76.9% amino acid identity also confer tolerance in transgenic Arabidopsis to lead [Pb(II)] and Cd(II), respectively.1,5,7 Since recombinant ACBP1 and ACBP2 bind linolenoyl-CoA and linoleoyl-CoA in vitro, they may possibly be involved in phospholipid repair in response to heavy metal stress at the plasma membrane.5,7 In contrast, ACBP3 is an extracellularly-localized protein8 while ACBP4, ACBP5 and ACBP6 are localized to cytosol.9,10 ACBP1 and ACBP6 have recently been shown to be involved in freezing stress.9,11 ACBP4 and ACBP5 bind oleoyl-CoA ester and their mRNA expressions are lightregulated.12,13 Besides acyl-CoA esters, some ACBPs also bind phospholipids.9,11,13 To investigate the biological function of ACBP2, we have proceeded to establish its interactors at the ankyrin repeats, including AtFP6,5 AtEBP6 and now lysoPL2 in the Plant Journal paper. While the significance in the interaction of ACBP2 with AtEBP awaits further investigations, some parallels can be drawn between those of ACBP2 with AtFP6 and with lysoPL2.  相似文献   

14.
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   

15.
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   

16.
The apical plasma membrane of young Arabidopsis root hairs has recently been found to contain a depolarisation-activated Ca2+ channel, in addition to one activated by hyperpolarisation. The depolarisation-activated Ca2+ channel may function in signalling but the possibility that the root hair apical plasma membrane voltage may oscillate between a hyperpolarized and depolarized state suggests a role in growth control. Plant NADPH oxidase activity has yet to be considered in models of oscillatory voltage or ionic flux despite its predicted electrogenicity and voltage dependence. Activity of root NADPH oxidase was found to be stimulated by restricting Ca2+ influx, suggesting that these enzymes are involved in sensing Ca2+ entry into cells.Key words: calcium, channel, NADPH oxidase, oscillation, root hairElevation of cytosolic free Ca2+ ([Ca2+]cyt) encodes plant cell signals.1 Reactive oxygen species (ROS) are potent regulators of the PM Ca2+ channels implicated in signalling and developmental increases in [Ca2+]cyt.1,2 Plasma membrane (PM) voltage (Vm) also plays a significant part in generating specific [Ca2+]cyt elevations through the opening of voltage-gated Ca2+-permeable channels, allowing Ca2+ influx.1,3 Patch clamp electrophysiological studies on the root hair apical PM of Arabidopsis have revealed co-localisation of hyperpolarisation-activated Ca2+ channels (HACCs),4 ROS-activated HACCs5 and depolarisation-activated Ca2+ channels (DACCs).6 The DACC characterisation pointed to the presence of a Cl-permeable conductance that was activated by moderate hyperpolarisation (−160 mV) but rapidly inactivated when the voltage was maintained at such negative values.6 This may be the R-type anion efflux conductance previously described in Arabidopsis root hair and root epidermal PM.7 Previous studies have shown that root hair PM also harbors K+ channels (mediating inward or outward flux)810 and a H+-ATPase.11 A key problem to address now is how these transporters interact to generate and be influenced by PM Vm, thus gating and in turn being regulated by their companion Ca2+ channels to encode developmental and environmental signals at the hair apex.A seminal study on the relationship between Vm and ionic fluxes in wheat root protoplasts not only confirmed oscillatory events but also determined that the PM can exist in three distinct states.12 In the “pump state” the H+-ATPase predominates, there is net H+ efflux and the hyperpolarized Vm is negative of the equilibrium potential for K+ (EK). In the “K state”, K+ permeability predominates but there is still net H+ efflux and Vm = EK. In the third state, there is net H+ influx and Vm > EK. In this depolarized H+-influx state, the H+-ATPase is thought to be inactive. Oscillations in PM Vm and H+ flux may be more profound in growing cells13,14 and oscillations between these states may explain the temporal changes in H+ flux recently observed at the apex of growing Arabidopsis root hairs.15 Peaks of H+ influx may reflect a depolarized Vm that could activate DACC, suggesting that DACC would play a significant role in growth regulation. The view has arisen that the HACC would be the main driver of growth, primarily because in patch clamp assays its current is greater than DACC46 and because resting Vm is usually found to be hyperpolarized. In a growing cell, with a Vm oscillating between a hyperpolarized and depolarized state, a DACC could just as well be a driver of growth given that the Ca2+ influx it permits could be amplified through intracellular release.The PM H+-ATPase traditionally lies at the core of models of voltage and ionic flux14,16 but in terms of [Ca2+]cyt regulation, the activity of PM NADPH oxidases must also now be considered. The Arabidopsis root hair apical PM also contains an NADPH oxidase (AtrbohC) that catalyses extracellular superoxide production.5 AtrbohC is implicated in the transition to polar growth at normal extracellular pH5 and also osmoregulation.17 NADPH oxidases catalyse the transport of electrons out of the cell and thus, in common with PM redox e efflux systems,18 their activity would depolarize the membrane voltage unless countered by cation efflux or anion influx.19 Two H+ would also be released into the cytosol for every NADPH used. The voltage-dependence of plant NADPH oxidases is unknown but e efflux by animal NADPH oxidases is fairly constant over negative Vm and decreases at very depolarized Vm.20 AtrbohC is implicated in generating oscillatory ROS at the root hair apex and loss of function affects magnitude and duration of apical H+ flux oscillations.15 The latter suggests that AtrbohC function does in some way affect Vm, a situation extending to other root cell types (such as the epidermis) expressing NADPH oxidases.21NADPH oxidase activity in roots is under developmental control but also responds to anoxia and nutrient deficiency22,23 to signal stress conditions. Blockade of PM Ca2+ channels by lanthanides increases superoxide production in tobacco suspension cells.24 This suggests that NADPH oxidases are involved in sensing the cell''s Ca2+ status and the prediction would be that extracellular Ca2+ chelation would increase their activity. To test this, superoxide anion production by excised Arabidopsis roots was measured using reduction of the tetrazolium dye XTT (Sodium, 3′-[1-[phenylamino-carbonyl]-3,4-tetrazolium]-bis(4-methoxy-6-nitro) benzene-sulphonic acid).25,26 Lowering extracellular Ca2+ from 0.5 mM to 1.4 µM by addition of 10 mM EGTA caused a mean 95% increase in diphenyliodinium-sensitive superoxide production (Fig. 1; n = 9), implicating NADPH oxidases as the source of this ROS. Stimulation of NADPH oxidase activity by decreasing Ca2+ influx at first appears contradictory as NADPH oxidases are stimulated by increased [Ca2+]cyt27 (Fig. 1). However, reduction of Ca2+ influx should promote voltage hyperpolarisation (just as block of K+ influx causes hyperpolarisation in root hairs28) and this could feasibly cause increased NADPH oxidase activity. Production of superoxide could then result in ROS-activated HACC activity5 to increase Ca2+ influx.Open in a separate windowFigure 1Superoxide anion production by Arabidopsis roots. Assay medium comprised 10 mM phosphate buffer with 0.5 mM CaCl2, 500 µM XTT, pH 6.0. Production was linear over the 30 min incubation period. Control, mean ± standard error, n = 9. Test additions were: 20 µM of the NADPH oxidase inhibitor diphenylene iodonium (DPI; n = 6); 100 µM of the Ca2+ ionophore A23187,30 to increase [Ca2+]cyt (n = 9); 10 mM of the chelator EGTA (n = 9). Dimethyl sulphoxide [DMSO; 1% (v/v)] was used as a carrier for XTT and DPI and a separate control for this is shown (n = 9).In addition to Vm, activities of PM transporters in vivo will be subject to other levels of regulation such as phosphorylation, nitrosylation and the action of [Ca2+]cyt itself. Distinct spatial separation of transporters will undoubtedly play a significant role in governing Vm and [Ca2+]cyt dynamics, particularly in growing cells. An NADPH oxidase has already been found sequestered in a potential PM microdomain in Medicago.29 While there is still much to do on the “inventory” of PM transporters involved in Ca2+ signalling in any given cell, placing them in context not only requires knowledge of their genetic identity but also modelling of their concerted action.  相似文献   

17.
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   

18.
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   

19.
Fetal cells migrate into the mother during pregnancy. Fetomaternal transfer probably occurs in all pregnancies and in humans the fetal cells can persist for decades. Microchimeric fetal cells are found in various maternal tissues and organs including blood, bone marrow, skin and liver. In mice, fetal cells have also been found in the brain. The fetal cells also appear to target sites of injury. Fetomaternal microchimerism may have important implications for the immune status of women, influencing autoimmunity and tolerance to transplants. Further understanding of the ability of fetal cells to cross both the placental and blood-brain barriers, to migrate into diverse tissues, and to differentiate into multiple cell types may also advance strategies for intravenous transplantation of stem cells for cytotherapeutic repair. Here we discuss hypotheses for how fetal cells cross the placental and blood-brain barriers and the persistence and distribution of fetal cells in the mother.Key Words: fetomaternal microchimerism, stem cells, progenitor cells, placental barrier, blood-brain barrier, adhesion, migrationMicrochimerism is the presence of a small population of genetically distinct and separately derived cells within an individual. This commonly occurs following transfusion or transplantation.13 Microchimerism can also occur between mother and fetus. Small numbers of cells traffic across the placenta during pregnancy. This exchange occurs both from the fetus to the mother (fetomaternal)47 and from the mother to the fetus.810 Similar exchange may also occur between monochorionic twins in utero.1113 There is increasing evidence that fetomaternal microchimerism persists lifelong in many child-bearing women.7,14 The significance of fetomaternal microchimerism remains unclear. It could be that fetomaternal microchimerism is an epiphenomenon of pregnancy. Alternatively, it could be a mechanism by which the fetus ensures maternal fitness in order to enhance its own chances of survival. In either case, the occurrence of pregnancy-acquired microchimerism in women may have implications for graft survival and autoimmunity. More detailed understanding of the biology of microchimeric fetal cells may also advance progress towards cytotherapeutic repair via intravenous transplantation of stem or progenitor cells.Trophoblasts were the first zygote-derived cell type found to cross into the mother. In 1893, Schmorl reported the appearance of trophoblasts in the maternal pulmonary vasculature.15 Later, trophoblasts were also observed in the maternal circulation.1620 Subsequently various other fetal cell types derived from fetal blood were also found in the maternal circulation.21,22 These fetal cell types included lymphocytes,23 erythroblasts or nucleated red blood cells,24,25 haematopoietic progenitors7,26,27 and putative mesenchymal progenitors.14,28 While it has been suggested that small numbers of fetal cells traffic across the placenta in every human pregnancy,2931 trophoblast release does not appear to occur in all pregnancies.32 Likewise, in mice, fetal cells have also been reported in maternal blood.33,34 In the mouse, fetomaternal transfer also appears to occur during all pregnancies.35  相似文献   

20.
Non-CG methylation is well characterized in plants where it appears to play a role in gene silencing and genomic imprinting. Although strong evidence for the presence of non-CG methylation in mammals has been available for some time, both its origin and function remain elusive. In this review we discuss available evidence on non-CG methylation in mammals in light of evidence suggesting that the human stem cell methylome contains significant levels of methylation outside the CG site.Key words: non-CG methylation, stem cells, Dnmt1, Dnmt3a, human methylomeIn plant cells non-CG sites are methylated de novo by Chromomethylase 3, DRM1 and DRM2. Chromomethylase 3, along with DRM1 and DRM2 combine in the maintenance of methylation at symmetric CpHpG as well as asymmetric DNA sites where they appear to prevent reactivation of transposons.1 DRM1 and DRM2 modify DNA de novo primarily at asymmetric CpH and CpHpH sequences targeted by siRNA.2Much less information is available on non-CG methylation in mammals. In fact, studies on mammalian non-CG methylation form a tiny fraction of those on CG methylation, even though data for cytosine methylation in other dinucleotides, CA, CT and CC, have been available since the late 1980s.3 Strong evidence for non-CG methylation was found by examining either exogenous DNA sequences, such as plasmid and viral integrants in mouse and human cell lines,4,5 or transposons and repetitive sequences such as the human L1 retrotransposon6 in a human embryonic fibroblast cell line. In the latter study, non-CG methylation observed in L1 was found to be consistent with the capacity of Dnmt1 to methylate slippage intermediates de novo.6Non-CG methylation has also been reported at origins of replication7,8 and a region of the human myogenic gene Myf3.9 The Myf3 gene is silenced in non-muscle cell lines but it is not methylated at CGs. Instead, it carries several methylated cytosines within the sequence CCTGG. Gene-specific non-CG methylation was also reported in a study of lymphoma and myeloma cell lines not expressing many B lineage-specific genes.10 The study focused on one specific gene, B29 and found heavy CG promoter methylation of that gene in most cell lines not expressing it. However, in two other cell lines where the gene was silenced, cytosine methylation was found almost exclusively at CCWGG sites. The authors provided evidence suggesting that CCWGG methylation was sufficient for silencing the B29 promoter and that methylated probes based on B29 sequences had unique gel shift patterns compared to non-methylated but otherwise identical sequences.10 The latter finding suggests that the presence of the non-CG methylation causes changes in the proteins able to bind the promoter, which could be mechanistically related to the silencing seen with this alternate methylation.Non-CG methylation is rarely seen in DNA isolated from cancer patients. However, the p16 promoter region was reported to contain both CG and non-CG methylation in breast tumor specimens but lacked methylation at these sites in normal breast tissue obtained at mammoplasty.11 Moreover, CWG methylation at the CCWGG sites in the calcitonin gene is not found in normal or leukemic lymphocyte DNA obtained from patients.12 Further, in DNA obtained from breast cancer patients, MspI sites that are refractory to digestion by MspI and thus candidates for CHG methylation were found to carry CpG methylation.13 Their resistance to MspI restriction was found to be caused by an unusual secondary structure in the DNA spanning the MspI site that prevents restriction.13 This latter observation suggests caution in interpreting EcoRII/BstNI or EcoRII/BstOI restriction differences as due to CWG methylation, since in contrast to the 37°C incubation temperature required for full EcoRII activity, BstNI and BstOI require incubation at 60°C for full activity where many secondary structures are unstable.The recent report by Lister et al.14 confirmed a much earlier report by Ramsahoye et al.15 suggesting that non-CG methylation is prevalent in mammalian stem cell lines. Nearest neighbor analysis was used to detect non-CG methylation in the earlier study on the mouse embryonic stem (ES) cell line,15 thus global methylation patterning was assessed. Lister et al.14 extend these findings to human stem cell lines at single-base resolution with whole-genome bisulfite sequencing. They report14 that the methylome of the human H1 stem cell line and the methylome of the induced pluripotent IMR90 (iPS) cell line are stippled with non-CG methylation while that of the human IMR90 fetal fibroblast cell line is not. While the results of the two studies are complementary, the human methylome study addresses locus specific non-CG methylation. Based on that data,14 one must conclude that non-CG methylation is not carefully maintained at a given site in the human H1 cell line. The average non-CG site is picked up as methylated in about 25% of the reads whereas the average CG methylation site is picked up in 92% of the reads. Moreover, non-CG methylation is not generally present on both strands and is concentrated in the body of actively transcribed genes.14Even so, the consistent finding that non-CG methylation appears to be confined to stem cell lines,14,15 raises the possibility that cancer stem cells16 carry non-CG methylation while their nonstem progeny in the tumor carry only CG methylation. Given the expected paucity of cancer stem cells in a tumor cell population, it is unlikely that bisulfite sequencing would detect non-CG methylation in DNA isolated from tumor cells since the stem cell population is expected to be only a very minor component of tumor DNA. Published sequences obtained by bisulfite sequencing generally report only CG methylation, and to the best of our knowledge bisulfite sequenced tumor DNA specimens have not reported non-CG methylation. On the other hand, when sequences from cell lines have been reported, bisulfite-mediated genomic sequencing8 or ligation mediated PCR17 methylcytosine signals outside the CG site have been observed. In a more recent study plasmid DNAs carrying the Bcl2-major breakpoint cluster18 or human breast cancer DNA13 treated with bisulfite under non-denaturing conditions, cytosines outside the CG side were only partially converted on only one strand18 or at a symmetrical CWG site.13 In the breast cancer DNA study the apparent CWG methylation was not detected when the DNA was fully denatured before bisulfite treatment.13In both stem cell studies, non-CG methylation was attributed to the Dnmt3a,14,15 a DNA methyltransferase with similarities to the plant DRM methyltransferase family19 and having the capacity to methylate non-CG sites when expressed in Drosophila melanogaster.15 DRM proteins however, possess a unique permuted domain structure found exclusively in plants19 and the associated RNA-directed non-CG DNA methylation has not been reproducibly observed in mammals despite considerable published2023 and unpublished efforts in that area. Moreover, reports where methylation was studied often infer methylation changes from 5AzaC reactivation studies24 or find that CG methylation seen in plants but not non-CG methylation is detected.21,22,25,26 In this regard, it is of interest that the level of non-CG methylation reported in stem cells corresponds to background non-CG methylation observed in vitro with human DNA methyltransferase I,27 and is consistent with the recent report that cultured stem cells are epigenetically unstable.28The function of non-CG methylation remains elusive. A role in gene expression has not been ruled out, as the studies above on Myf3 and B29 suggest.9,10 However, transgene expression of the bacterial methyltransferase M.EcoRII in a human cell line (HK293), did not affect the CG methylation state at the APC and SerpinB5 genes29 even though the promoters were symmetrically de novo methylated at mCWGs within each CCWGG sequence in each promoter. This demonstrated that CG and non-CG methylation are not mutually exclusive as had been suggested by earlier reports.9,10 That observation is now extended to the human stem cell line methylome where CG and non-CG methylation co-exist.14 Gene expression at the APC locus was likewise unaffected by transgene expression of M.EcoRII. In those experiments genome wide methylation of the CCWGG site was detected by restriction analysis and bisulfite sequencing,29 however stem cell characteristics were not studied.Many alternative functions can be envisioned for non-CG methylation, but the existing data now constrains them to functions that involve low levels of methylation that are primarily asymmetric. Moreover, inheritance of such methylation patterns requires low fidelity methylation. If methylation were maintained with high fidelity at particular CHG sites one would expect that the spontaneous deamination of 5-methylcytosine would diminish the number of such sites, so as to confine the remaining sites to those positions performing an essential function, as is seen in CG methylation.3033 However, depletion of CWG sites is not observed in the human genome.34 Since CWG sites account for only about 50% of the non-CG methylation observed in the stem cell methylome14 where methylated non-CG sites carry only about 25% methylation, the probability of deamination would be about 13% of that for CWG sites that are subject to maintenance methylation in the germ line. Since mutational depletion of methylated cytosines has to have its primary effect on the germ line, if the maintenance of non-CG methylation were more accurate and more widespread, one would have had to argue that stem cells in the human germ lines lack CWG methylation. As it is the data suggests that whatever function non-CG methylation may have in stem cells, it does not involve accurate somatic inheritance in the germ line.The extensive detail on non-CG methylation in the H1 methylome14 raises interesting questions about the nature of this form of methylation in human cell lines. A key finding in this report is the contrast between the presence of non-CG methylation in the H1 stem cell line and its absence in the IMR90 human fetal lung fibroblast cell line.14 This suggests that it may have a role in the origin and maintenance of the pluripotent lineage.14By analogy with the well known methylated DNA binding proteins specific for CG methylation,35 methylated DNA binding proteins that selectively bind sites of non-CG methylation are expected to exist in stem cells. Currently the only protein reported to have this binding specificity is human Dnmt1.3638 While Dnmt1 has been proposed to function stoichiometrically39 and could serve a non-CG binding role in stem cells, this possibility and the possibility that other stem-cell specific non-CG binding proteins might exist remain to be been explored.Finally, the nature of the non-CG methylation patterns in human stem cell lines present potentially difficult technical problems in methylation analysis. First, based on the data in the H1 stem cell methylome,40 a standard MS-qPCR for non-CG methylation would be impractical because non-CG sites are infrequent, rarely clustered and are generally characterized by partial asymmetric methylation. This means that a PCR primer that senses the 3 adjacent methylation sites usually recommended for MS-qPCR primer design41,42 cannot be reliably found. For example in the region near Oct4 (Chr6:31,246,431), a potential MS-qPCR site exists with a suboptimal set of two adjacent CHG sites both methylated on the + strand at Chr6:31,252,225 and 31,252,237.14,40 However these sites were methylated only in 13/45 and 30/52 reads. Thus the probability that they would both be methylated on the same strand is about 17%. Moreover, reverse primer locations containing non-CG methylation sites are generally too far away for practical bisulfite mediated PCR. Considering the losses associated with bisulfite mediated PCR43 the likelihood that such an MS-qPCR system would detect non-CG methylation in the H1 cell line or stem cells present in a cancer stem cell niche44,45 is very low.The second difficulty is that methods based on the specificity of MeCP2 and similar methylated DNA binding proteins for enriching methylated DNA (e.g., MIRA,46 COMPARE-MS47) will discard sequences containing non-CG methylation since they require cooperative binding afforded by runs of adjacent methylated CG sites for DNA capture. This latter property of the methylated cytosine capture techniques makes it also unlikely that methods based on 5-methylcytosine antibodies (e.g., meDIP48) will capture non-CG methylation patterns accurately since the stem cell methylome shows that adjacent methylated non-CG sites are rare in comparison to methylated CG sites.14In summary, whether or not mammalian stem cells in general or human stem cells in particular possess functional plant-like methylation patterns is likely to continue to be an interesting and challenging question. At this point we can conclude that the non-CG patterns reported in human cells appear to differ significantly from the non-CG patterns seen in plants, suggesting that they do not have a common origin or function.  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