Summary
Prions are infectious proteins, without the need for an accompanying nucleic acid. Nonetheless, there are connections of prions with translation and RNA, which we explore here. Most prions are based on self-propagating amyloids. The yeast [PSI+] prion is an amyloid of Sup35p, a subunit of the translation termination factor. The normal function of the Sup35p prion domain is in shortening the 3' polyA of mRNAs and thus in mRNA turnover. The [ISP+] prion is so named because it produces anti-suppression, the opposite of the effect of [PSI+]. Another connection of prions with translation is the influence on prion propagation and generation of ribosome-associated chaperones, the Ssb's, and a chaperone activity intrinsic to the 60S ribosomal subunits.
Keywords: prion, amyloid, parallel in-register, chaperones, Sup35, mRNA turnover, translation termination
Prions are infectious proteins that do not require an accompanying nucleic acid to transmit a disease or a trait. Prions of mammals include the uniformly fatal transmissible spongiform encephalopathies, such as scrapie of sheep, Creutzfeldt-Jakob disease of humans, bovine spongiform encephalopathy and chronic wasting disease of elk and deer. These diseases are amyloidoses of PrP, a cell surface protein of uncertain function (reviewed in (1, 2)). The recognition that [URE3] and [PSI+], two long-known cytoplasmic genetic elements of S. cerevisiae (3, 4), are prions of Ure2p and Sup35p, respectively (5), accelerated progress in the prion field. Yeast and filamentous fungi are now known to have a number of amyloid - based prions and two prions based on enzymatic processes (Table I).
Table 1.
Prions
| Prion | Organism | Protein | Protein function or phenotype | Ref. | 
|---|---|---|---|---|
| Amyloid-based prions | ||||
| TSEs | mammals | PrP | non-essential, function unclear | (1, 2) | 
| [URE3] | S. cerevisiae | Ure2p | nitrogen catabolism regulator | (5) | 
| [PSI+] | S. cerevisiae | Sup35p | translation termination factor | (5) | 
| [PIN+] | S. cerevisiae | Rnq1 | rare priming of other prions | (68) | 
| [SWI+] | S. cerevisiae | Swi1p | transcription factor | (69) | 
| [MCA] | S. cerevisiae | Mca1 | metacaspase homolog | (70) | 
| [OCT] | S. cerevisiae | Cyc8p | transcriptional co-repressor | (71) | 
| [Het-s] | P. anserina | HET-s | heterokaryon incompatibility | (72) | 
| [ISP+] | S. cerevisiae | ?Sfp1 | antisuppression | (47) | 
| [MOT] | S. cerevisiae | Mot3p | transcription factor | (73) | 
| Enzyme activity-based prions | ||||
| [BETA] | S. cerevisiae | Prb1 | active vacuolar protease B | (7) | 
| [C] | P. anserina | MAP kinases | 'crippled growth' | (8) | 
Yeast prions can determine phenotypic traits which, in many cases, were the basis of their discovery. Amyloid formation by the Ure2 protein makes it inefficient at repressing the genes encoding enzymes and transporters needed for using poor nitrogen sources when a good nitrogen source is available. Derepression of DAL5, encoding the allantoate transporter, is used to detect the [URE3] prion. Sup35p is a subunit of the translation termination factor, and its amyloid formation in [PSI+] cells makes it inefficient resulting in increased readthrough of termination codons.
Not all prions involve amyloid. The [β] prion of S. cerevisiae is simply the active form of the vacuolar protease B, which, in the absence of protease A, can activate (and is needed to activate) its own precursor (6, 7). Cells without the active protease B remain so, and those with it propagate this activity by converting the inactive precursor to active form. The active form is infectious, and this system has all the genetic properties expected for a prion. The [C] trait of Podospora appears to be similar, possibly involving self-phosphorylating MAP kinases as the basis of the trait (8).
In this review, we briefly survey the growing menagerie of prions, and explore the connections of prions with translation and RNA. Although the fundemental prion phenomenon is one of protein-protein interaction, there are several interesting interactions of the translation process and prions.
Structural basis of yeast prions
Each prion protein has a prion domain, a limited part of the protein which is sufficient to transmit the prion in vivo, and whose amyloid form made from recombinant protein can infect yeast cells with the prion (9–14). In amyloid formed by the full length Ure2p, for example, the prion domain has changed from being unstructured to the highly structured amyloid form, while the rest of the molecule is essentially unchanged (15–18). Amyloid formed by the prion domains of Ure2p, Sup35p and Rnq1p each have an in-register parallel beta sheet architecture with folds of the sheet along the filament axis (Fig. 1)(19–21). Thus, the surface of the prion filaments have strips of aligned amino acid residues: a line of glutamines or asparagines can form the "polar zipper" - a line of hydrogen bonds between the amide side chains that stabilize the structure (22–24), a line of serine or threonine residues can form pairwise hydrogen bonds, a line of a hydrophobic residue will likewise have positive interactions. It is suggested that variation in the location of the folds and the extent of the beta sheet may be the points of difference among prion variants.
Fig. 1. In-register parallel β-sheet structure explains how prions transmit their conformation to the normal form of the protein.
The unstructured prion domain of monomers are forced to assume the conformation of the protein in the amyloid filaments by interactions with the main chain and side chain parts of the identical residues of the protein.
This structure suggests a mechanism by which information is inherited by a protein – how a protein can act as a gene (25). The prion domains are unstructured, and acquire the structure of the molecules in the amyloid filament as they add on to the end. The favorable interactions between adjacent identical residues on the last molecule that is part of the filament and the newly joining molecule - the polar zippers, other hydrogen bonds along the filament, and hydrophobic interactions – direct the new molecule to change from being unstructured to having the structure of the molecules in the filament. Segments forming beta strands in the filament molecules will be beta strands in the new molecule; turns in the filament will be turns in the new molecule (Fig. 1)(25). Thus the prion form templates the conformation of new prion protein monomers joining the filament, just as parental DNA strands template the sequence of new DNA strands.
Biological roles of yeast and fungal prions
While mammalian prions are uniformly fatal, yeast and fungal prions are compatible with continued growth, and it has been suggested that some may be advantageous to their hosts (26–28). Plate tests of growth under various conditions cannot tell whether prions are beneficial or detrimental without knowledge of the distribution of the conditions tested in the ecological niche of yeast (29). It was suggested that [PSI+] was a general advantage under stress conditions, or at least at high temperature or high ethanol conditions (27). In another report, the effect of [PSI+] on tolerance to ethanol, heat or other stresses varied with yeast strains (28). A third group, using the same strains as the second group, found mostly different results (30).
How then can one determine if a prion is beneficial or detrimental? Deleterious viruses, plasmids, bacteria and prions (e.g. scrapie and chronic wasting disease) are easily found in the wild. Their infectivity allows them to spread faster than the damage they cause to their hosts selects against their abundance. Certainly an infectious element that is advantageous would be everywhere. In fact, the [URE3] and [PSI+] prions are not found in any of the wild strains surveyed (31), indicating that these prions are diseases of yeast. Whether transient variants of yeast prions could have beneficial phenotypes has not been reported. In contrast, 80% of wild strains carry the [Het-s] prion, suggesting that it is a benefit to its host (32). Indeed, the heterokaryon incompatibility function provided by the [Het-s] prion is known to be advantageous to the host. However, [Het-s] is the basis for a meiotic drive phenomenon (where a gene promotes its own inheritance by disabling gametes with other alleles, rather than benefiting its host), so that even in this case, the prion may be a disadvantage (32).
While rather few species have been examined, the ability of the N-terminal domains of Ure2p from other species to support prion formation seems to be sporadic (33, 34)[H.E. Edskes, A. Engel & R. Wickner, in preparation], suggesting that prion formation is not preserved in evolution, but is rather the price yeast pays for another function. In fact, the prion domain prevents the rapid turnover of the Ure2 protein and so is necessary for efficient nitrogen catabolite repression (35). Similarly 25% of wild strains of S. cerevisiae examined had a large deletion in their Sup35 prion domains preventing them from becoming [PSI+], indicating that prion formation in this case as well is not conserved even within cerevisiae (36). The Sup35p prion domain also has a function that is unrelated to prion formation, being necessary for the normal process of mRNA turnover through a role in shortening of the mRNA's 3' polyA (37) (see next section).
Sup35p roles in translation termination, mRNA turnover and prion formation
Sup35p/eRF3 is, with Sup45p/eRF1, a subunit of the translation termination factor. Sup45p recognizes stop codons and activates the ribosome's cleavage of the peptide-tRNA bond. Sup35p, through its GTPase activity, stimulates Sup45p (38–40). Prion formation by Sup35p (the [PSI+] prion) is detected by increased readthrough of premature termination codons (3). One example of this pathology is elevated readthrough of the "shifty stop" codon controlling antizyme, a protein that targets ornithine decarboxylase for degradation and thereby controls polyamine biosynthesis (30). While deletion of the Sup35p N-terminal prion domain does not affect the efficiency of translation termination (as judged by readthrough of premature termination codons) (41), the presence of the prion domain does enhance readthrough in the context of a number of mutations in the C-terminal domain (42).
The Sup35p N-terminal prion domain also interacts specficially with Pab1p, the polyA binding protein (43, 44). Deletion of the Sup35 prion domain lengthens the half-life of the various mRNAs tested, both native yeast messages and foreign mRNAs (37). This mRNA lifetime expansion is accompanied by an increase in the mean mRNA polyA length, indicating that the Sup35p prion domain has a role in the polyA shortening process that preceeds, and is a signal for mRNA decay. This interaction is not specific to the S. cerevisiae Sup35p N-terminal domain, and was actually first documented for the human Gspt/Sup35 N-terminal domain (43). These results indicate that the Sup35 prion domain is not simply there to enable the cell to form prions. Rather, the Sup35p prion domain has a normal function in mRNA turnover through its role in 3' polyA shortening.
[ISP+] may be a prion of Sfp1p, a regulator of ribosome biogenesis
[ISP+] is a non-chromosomal genetic element that gets its name from having the opposite effect to that of [PSI+], namely it confers antisuppression (45). As discussed above, Sup35p is a subunit of the translation termination factor, and thus mutations of SUP35 have a suppressor effect, that is, a termination codon is not promptly recognized as such because of a deficiency of normal termination factor, and there is more time for a mistranslation event to occur, such as the pairing of a suppressor tRNA to the stop codon. The sup35-10 and sup35-25 missense mutants have such a suppressor effect with the nonsense mutants, his7-1 and lys2-87. Volkov et al. noted that after culturing strains containing these suppressible mutations and sup35-10 or sup35-25, they often found that the strains had lost the suppressor effect of the sup35 mutation. This antisuppressor effect was frequently eliminated by growth in the presence of guanidine HCl, but unlike curing of [PSI+] or [URE3], concentrations of > 5 mM were needed to observe curing, and the efficiency of curing was rather low (45). From a cured strain ([isp−]), it was possible to again isolate [ISP+] derivatives. Such “reversible curing” is one of the classical genetic criteria for identifying a prion (5).
On crossing an [ISP+] antisuppressor – carrying strain with an [isp−] (guanidine-cured) strain (both carrying the sup35-10 or sup35-25 mutation and the suppressible his7-1 and lys2-87 alleles), the meiotic spore clones showed 4antisuppressed : 0 segregation, and each of these antisuppressed spore clones could be cured of the antisuppression by growth in guanidine (45). [ISP+] could also be transferred to an [isp−] strain by cytoduction (mating followed by cytoplasmic mixing but without nuclear fusion), but only 16% of cytoductants received the [ISP+] element (45). Guanidine curing of [PSI+] has been shown to be due to the inhibition of Hsp104 (46), but [ISP+] was not cured by either deletion of HSP104 or by its overexpression (45). Although the [ISP+] genetic element is manifested through its effects on sup35 mutants, propagation of [ISP+] does not depend on the Sup35p prion domain, nor does overexpression of Sup35p increase the frequency of [ISP+] appearance in an initially [isp−] strain (45).
Recently, it has been shown that propagation of [ISP+] depends on the Sfp1 protein (47), a transcription factor controlling ribosome biogenesis (48). "Thus [ISP+] is apparently a prion of Sfp1p, and the involvement of Sfp1p in controlling ribosome biogenesis implies that it could affect suppression efficiency."
Mammalian prions, translation and RNA
Propagation of the PrP-based transmissible spongiform encephalopathies involves conversion of PrPC, a cell - surface, glycophosphatidylinositol-anchored protein, to PrPSc, an amyloid form of the protein. This conversion process occurs on the cell surface, or in endosomes, long after the translation process (49). An in vitro prion propagation system has been developed, reminiscent of the polymerase chain reaction, in which PrPSc serves as template for polymerization of amyloid starting with brain extracts containing PrPC. In this system, RNA stimulates the reaction, but this effect is not sequence specific, and even polyA can serve this purpose (50).
An internal ribosome initiation site in URE2 can affect [URE3]
The prion domain of Ure2p extends from about residues 1 to 65 or 70 as judged by the part that is sufficient for prion induction (10) or the portion that becomes protease - resistant in amyloid of Ure2p (16). However, the Q/N rich property extends to about residue 89 and the structured C-terminal domain begins at about residue 97 (51)(52). Ure2p begins with a pair of methionine residues, and the third methionine occurs at residue 94.
Komar et al. have shown that initiation occurs at the residue 94 with a significant frequency (53). The Ure2p94 – 354 fragment represented about 20% of Ure2 protein in extracts of wild-type strains, but changing of codon 94 from AUG to CTT eliminated this fragment, indicating that it was not due to translation of a minor transcript or to degradation of the unstructured prion domain (53). Inserting URE2 ORF nt 203–368 (encoding amino acids 68–122) into a bicistronic vector confirmed that URE2 contains an internal ribosome entry site (IRES) promoting cap-independent translation. Because Ure2p lacking the prion domain is more rapidly degraded (35), the real frequency of initiation at Met94 may be higher than it appears.
Overproduction of the C-terminal domain cures [URE3] (54). Replacing the normal cap-binding protein, eIF4E, with a temperature sensitive version, increased the fraction of Ure2p starting at Met94 to near 50% and destabilized the [URE3] prion. In addition, the 94AUG to CTT mutation results in the high frequency spontaneous development of the [URE3] prion (53). These results suggest that the presence of the URE2 IRES keeps down the frequency of [URE3] generation.
While the data in Komar et al. seems clear, it should be mentioned that Shewmaker et al. did not detect the 30 kDa Ure2 species in normal cells using a genomic copy of Ure2p with a C-terminal HA epitope tag (35). Moreover, we do not observe high frequency [URE3] generation or the [URE3] prion phenotype in the URE2 94AUG->CTT mutant (unpublished results).
Ribosome-associated chaperones in prion generation and propagation
The Hsp70 superfamily of chaperones includes the Ssb group, in yeast comprising two nearly identical proteins largely associated with ribosomes in polysomes, and believed to facilitate folding of nascent polypeptides (55). Chernoff et al. have shown that Ssb1p and Ssb2p have antagonistic activities toward propagation and generation of [PSI+](56). Overproduction of the disaggregase Hsp104 cures [PSI+] and concomitant overproduction of Ssb1p increases the curing efficiency. In fact, overproduction of Ssb1 alone can cure [PSI+] (57). The ssb1Δ ssb2Δ double mutant also shows a considerably increased frequency with which [PSI+] arises de novo. (56). These results suggest the possibility that nascent (ribosome-associated) Sup35p may be able to associate with the growing amyloid filaments, but there is no direct evidence for this. Note that although the Ssb proteins are ribosome-associated, conversion of Sup35p to the prion form need not occur co-translationally (58).
Another ribosome-associated chaperone activity is an activity of the large subunit rRNA itself (59). For example, the E. coli 23S rRNA can renature lactate dehydrogenase in a reaction inhibited by the ribosome-active antibiotics chloramphenicol and erythromycin.
Blondel embarked on a search for treatments for mammalian prions by screening compounds for curing of the yeast prions [PSI+] and [URE3], and then testing promising candidates in scrapie-infected tissue culture cells and mice (60). Remarkably, he found that several compounds active against yeast also cleared tissue culture cells of scrapie infection and prolonged survival of scrapie-infected mice (61). Affinity columns made with two drugs identified in this screen, 6-aminophenanthridine (6-AP) and guanabenz (GA), each bound ribosomes, while similar drugs inactive against prions did not. While neither drug affected translation in vivo, both inhibited the ribosome - associated chaperone activity of S. cerevisiae ribosomes, measured by its ability to renature carbonic anhydrase (62). This work is important both in providing a potential avenue toward treatment of the thus far intractable prion diseases, and a new approach to the incompletely characterized ribosomal chaperone activity.
While the intricate interactions of many chaperones of various classes with prions (reviewed in (63)) make it perhaps no surprise that a ribosomal chaperone activity should affect prion propagation, it is not yet clear whether this activity is important during the synthesis of the prion proteins, or is acting independent of translation. Moreover, the current evidence does not unequivocally rule out the possibility that the prion - blocking and rRNA chaperone inhibition activities of these drugs are separate actions of a single compound.
Cytoplasmic polyadenylation factor is speculated to be a prion in yeast
The marine snail Aplysia is used in studies of learning because of its large neurons and relatively simple nervous system (64). Translation is believed to play a role in synaptic plasticity, the changes in synapses consequent to their repeated stimulation that are believed to have a role in memory. The cytoplasmic polyadenylation element binding factor (CPEB) regulates mRNA polyadenylation, and thus translation, in many vertebrates and invertebrates by binding to a 3' UTR site called CPE along with Maskin, Symplekin, CPSF and other factors required for the regulation (65). When CPEB is activated by specific kinases, it induces polyadenylation of the mRNA to which it is bound by a cytoplasmic polyA polymerase. CPEB appears to be involved in mediating synaptic plasticity (65). The presence of a Gln/Asn-rich sequence in the N-terminal part of Aplysia CPEB, resembling yeast prion domains, led to speculation that prion formation by CPEB might have a role in memory.
To test this speculation, the ability of CPEB to act as a prion in S. cerevisiae was examined (66). The CPE site was placed in the 3' UTR of lacZ in yeast, and it was reported that this construct was only expressed when CPEB was also expressed and was in an aggregated form, and that mRNA levels were independent of CPEB expression or aggregation in this yeast system (66). However, there are no yeast homologs of Maskin or Symplekin, nor does yeast have a cytoplasmic polyA polymerase, so it is unlikely that CPEB could have its normal function in yeast. While non-polyA mRNAs are stable in Xenopus oocytes in which many of the original CPEB mechanism studies were done, yeast mRNAs lacking polyA are rapidly degraded (e.g., (67)). While aggregation of CPEB was observed in some cells and not in others, the evidence that this was a prion phenomenon was incomplete. For example, does overproduction of CPEB increase the frequency with which the putative prion arises? Moreover, deletion of the Q/N-rich putative prion domain did not affect the activation of CPEB. It is also paradoxical that aggregation was necessary for CPEB action. One would expect that, as in the case of Sup35p, aggregation would impair ability to reach the mRNAs where the action must occur. Clearly, further work will be necessary to establish whether CPEB can act as a prion in yeast.
Conclusions
While the prion phenomenon is primarily one of protein misfolding, it is clear that there are several connections with RNA and translation. Two prions primarily affect translation, and several have effects on transcription, but of course a prion could, in principle, affect any process a protein can affect. The translation – associated chaperones, both the intrinsic ribosomal chaperone activity and the ribosome-associated Ssb Hsp70s, have important effects on prion generation and propagation. The detailed mechanisms of each will doubtless be of increasing interest and there is a prospect of prion therapies based on drugs affecting ribosomal chaperones. Since the folding of proteins during and immediately after synthesis is a dicey process, it seems inevitable that more connections will be found between this process and the misfolding of prion proteins.
Fig. 2. Prions and translation.
The Sup35 prion protein is a component of the translation termination factor, and is necessary for the shortening of the 3' polyA of mRNAs that is a stage in mRNA degradation. The ribosome-associated Ssb1 chaperone protein and the chaperone activity of the 60S ribosome subunits also affect prion generation and propagation. The [ISP+] prion also affects translation by an as yet unknown mechanism.
Literature cited
- 1.Caughey B, editor. Prion Proteins. San Diego: Academic Press; 2001. [Google Scholar]
 - 2.Prusiner SB, editor. Prion Biology and Diseases. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press; 2004. [Google Scholar]
 - 3.Cox BS. PSI, a cytoplasmic suppressor of super-suppressor in yeast. Heredity. 1965;20:505–521. [Google Scholar]
 - 4.Lacroute F. Non-Mendelian mutation allowing ureidosuccinic acid uptake in yeast. J. Bacteriol. 1971;106:519–522. doi: 10.1128/jb.106.2.519-522.1971. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 5.Wickner RB. [URE3] as an altered URE2 protein: evidence for a prion analog in S. cerevisiae. Science. 1994;264:566–569. doi: 10.1126/science.7909170. [DOI] [PubMed] [Google Scholar]
 - 6.Zubenko GS, Park FJ, Jones EW. Genetic properties of mutations at the PEP4 locus in Saccharomyces cerevisiae. Genetics. 1982;102:679–690. doi: 10.1093/genetics/102.4.679. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 7.Roberts BT, Wickner RB. A class of prions that propagate via covalent auto-activation. Genes Dev. 2003;17:2083–2087. doi: 10.1101/gad.1115803. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 8.Kicka S, Bonnet C, Sobering AK, Ganesan LP, Silar P. A mitotically inheritable unit containing a MAP kinase module. Proc. Natl. Acad. Sci. USA. 2006;103:13445–13450. doi: 10.1073/pnas.0603693103. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 9.TerAvanesyan A, Dagkesamanskaya AR, Kushnirov VV, Smirnov VN. The SUP35 omnipotent suppressor gene is involved in the maintenance of the non-Mendelian determinant [psi+] in the yeast Saccharomyces cerevisiae. Genetics. 1994;137:671–676. doi: 10.1093/genetics/137.3.671. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 10.Masison DC, Wickner RB. Prion-inducing domain of yeast Ure2p and protease resistance of Ure2p in prion-containing cells. Science. 1995;270:93–95. doi: 10.1126/science.270.5233.93. [DOI] [PubMed] [Google Scholar]
 - 11.Masison DC, Maddelein M-L, Wickner RB. The prion model for [URE3] of yeast: spontaneous generation and requirements for propagation. Proc. Natl. Acad. Sci. USA. 1997;94:12503–12508. doi: 10.1073/pnas.94.23.12503. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 12.Vitrenko YA, Pavon ME, Stone SI, Liebman SW. Propagation of the [PIN+] prion by fragments of Rnq1 fused to GFP. Curr. Genet. 2007;51:309–319. doi: 10.1007/s00294-007-0127-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 13.King CY, Diaz-Avalos R. Protein-only transmission of three yeast prion strains. Nature. 2004;428:319–323. doi: 10.1038/nature02391. [DOI] [PubMed] [Google Scholar]
 - 14.Tanaka M, Chien P, Naber N, Cooke R, Weissman JS. Conformational variations in an infectious protein determine prion strain differences. Nature. 2004;428:323–328. doi: 10.1038/nature02392. [DOI] [PubMed] [Google Scholar]
 - 15.Taylor KL, Cheng N, Williams RW, Steven AC, Wickner RB. Prion domain initiation of amyloid formation in vitro from native Ure2p. Science. 1999;283:1339–1343. doi: 10.1126/science.283.5406.1339. [DOI] [PubMed] [Google Scholar]
 - 16.Baxa U, Taylor KL, Wall JS, Simon MN, Cheng N, Wickner RB, Steven A. Architecture of Ure2p prion filaments: the N-terminal domain forms a central core fiber. J. Biol. Chem. 2003;278:43717–43727. doi: 10.1074/jbc.M306004200. [DOI] [PubMed] [Google Scholar]
 - 17.Pierce MM, Baxa U, Steven AC, Bax A, Wickner RB. Is the prion domain of soluble Ure2p unstructured? Biochemistry. 2005;44:321–328. doi: 10.1021/bi047964d. [DOI] [PubMed] [Google Scholar]
 - 18.Baxa U, Cheng N, Winkler DC, Chiu TK, Davies DR, Sharma D, Inouye H, Kirschner DA, Wickner RB, Steven AC. Filaments of the Ure2p prion protein have a cross-beta core structure. J Struct Biol. 2005;150:170–179. doi: 10.1016/j.jsb.2005.02.007. [DOI] [PubMed] [Google Scholar]
 - 19.Shewmaker F, Wickner RB, Tycko R. Amyloid of the prion domain of Sup35p has an in-register parallel β-sheet structure. Proc. Natl. Acad. Sci. USA. 2006;103:19754–19759. doi: 10.1073/pnas.0609638103. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 20.Baxa U, Wickner RB, Steven AC, Anderson D, Marekov L, Yau W-M, Tycko R. Characterization of β-sheet structure in Ure2p1-89 yeast prion fibrils by solid state nuclear magnetic resonance. Biochemistry. 2007;46:13149–13162. doi: 10.1021/bi700826b. [DOI] [PubMed] [Google Scholar]
 - 21.Wickner RB, Dyda F, Tycko R. Amyloid of Rnq1p, the basis of the [PIN+] prion, has a parallel in-register β-sheet structure. Proc Natl Acad Sci U S A. 2008;105:2403–2408. doi: 10.1073/pnas.0712032105. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 22.Perutz MF, Johnson T, Suzuki M, Finch JT. Glutamine repeats as polar zippers: their possible role in inherited neurodegenerative diseases. Proc. Natl. Acad. Sci. USA. 1994;91:5355–5358. doi: 10.1073/pnas.91.12.5355. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 23.Chan JCC, Oyler NA, Yau W-M, Tycko R. Parallel β-sheets and polar zippers in amyloid fibrils formed by residues 10--39 of the yeast prion protein Ure2p. Biochemistry. 2005;44:10669–10680. doi: 10.1021/bi050724t. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 24.Nelson R, Sawaya MR, Balbirnie M, Madsen AO, Riekel C, Grothe R, Eisenberg D. Structure of the cross-β spine of amyloid-like fibrils. Nature. 2005;435:773–778. doi: 10.1038/nature03680. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 25.Wickner RB, Shewmaker F, Kryndushkin D, Edskes HK. Protein inheritance (prions) based on parallel in-register β-sheet amyloid structures. Bioessays. 2008;30:955–964. doi: 10.1002/bies.20821. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 26.Wickner RB. A new prion controls fungal cell fusion incompatibility. Proc. Natl. Acad. Sci. USA. 1997;94:10012–10014. doi: 10.1073/pnas.94.19.10012. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 27.Eaglestone SS, Cox BS, Tuite MF. Translation termination efficiency can be regulated in Saccharomyces cerevisiae by environmental stress through a prion-mediated mechanism. EMBO J. 1999;18:1974–1981. doi: 10.1093/emboj/18.7.1974. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 28.True HL, Lindquist SL. A yeast prion provides a mechanism for genetic variation and phenotypic diversity. Nature. 2000;407:477–483. doi: 10.1038/35035005. [DOI] [PubMed] [Google Scholar]
 - 29.Partridge L, Barton NH. Evolving evolvability. Nature. 2000;407:457–458. doi: 10.1038/35035173. [DOI] [PubMed] [Google Scholar]
 - 30.Namy O, Galopier A, Martini C, Matsufuji S, Fabret C, Rousset C. Epigenetic control of polyamines by the prion [PSI+] Nat. Cell. Biol. 2008 doi: 10.1038/ncb1766. Epub ahead of print. [DOI] [PubMed] [Google Scholar]
 - 31.Nakayashiki T, Kurtzman CP, Edskes HK, Wickner RB. Yeast prions [URE3] and [PSI+] are diseases. Proc Natl Acad Sci U S A. 2005;102:10575–10580. doi: 10.1073/pnas.0504882102. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 32.Dalstra HJP, Swart K, Debets AJM, Saupe SJ, Hoekstra RF. Sexual transmission of the [Het-s] prion leads to meiotic drive in Podospora anserina. Proc Natl Acad Sci U S A. 2003;100:6616–6621. doi: 10.1073/pnas.1030058100. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 33.Talarek N, Maillet L, Cullin C, Aigle M. The [URE3] prion is not conserved among Saccharomyces species. Genetics. 2005;171:23–54. doi: 10.1534/genetics.105.043489. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 34.Edskes HK, McCann LM, Hebert AM, Wickner RB. Prion variants and species barriers among Saccharomyces Ure2 proteins. Genetics. 2009;181:1159–1167. doi: 10.1534/genetics.108.099929. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 35.Shewmaker F, Mull L, Nakayashiki T, Masison DC, Wickner RB. Ure2p function is enhanced by its prion domain in Saccharomyces cerevisiae. Genetics. 2007;176:1557–1565. doi: 10.1534/genetics.107.074153. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 36.Resende CG, Outeiro TF, Sands L, Lindquist S, Tuite MF. Prion protein gene polymorphisms in Saccharomyces cerevisiae. Mol. Microbiol. 2003;49:1005–1017. doi: 10.1046/j.1365-2958.2003.03608.x. [DOI] [PubMed] [Google Scholar]
 - 37.Hosoda N, Kobayashii T, Uchida N, Funakoshi Y, Kikuchi Y, Hoshino S, Katada T. Translation termination factor eRF3 mediates mRNA decay through the regulation of deadenylation. J. Biol. Chem. 2003;278:38287–38291. doi: 10.1074/jbc.C300300200. [DOI] [PubMed] [Google Scholar]
 - 38.Frolova L, LeGoff X, Rasmussen HH, Cheperegin S, Drugeon G, Kress M, Arman I, Haenni A-L, Celis JE, Philippe M, Justesen J, Kisselev L. A highly conserved eukaryotic protein family possessing properties of polypeptide chain release factor. Nature. 1994;372:701–703. doi: 10.1038/372701a0. [DOI] [PubMed] [Google Scholar]
 - 39.Stansfield I, Jones KM, Kushnirov VV, Dagkesamanskaya AR, Poznyakovski AI, Paushkin SV, Nierras CR, Cox BS, Ter-Avanesyan MD, Tuite MF. The products of the SUP45 (eRF1) and SUP35 genes interact to mediate translation termination in Saccharomyces cerevisiae. EMBO J. 1995;14:4365–4373. doi: 10.1002/j.1460-2075.1995.tb00111.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 40.Zhouravleva G, Frolova L, LeGoff X, LeGuellec R, Inge-Vectomov S, Kisselev L, Philippe M. Termination of translation in eukaryotes is governed by two interacting polypeptide chain release factors, eRF1 and eRF3. EMBO J. 1995;14:4065–4072. doi: 10.1002/j.1460-2075.1995.tb00078.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 41.TerAvanesyan MD, Kushnirov VV, Dagkesamanskaya AR, Didichenko SA, Chernoff YO, Inge-Vechtomov SG, Smirnov VN. Deletion analysis of the SUP35 gene of the yeast Saccharomyces cerevisiae reveals two non-overlapping functional regions in the encoded protein. Mol. Microbiol. 1993;7:683–692. doi: 10.1111/j.1365-2958.1993.tb01159.x. [DOI] [PubMed] [Google Scholar]
 - 42.Volkov K, Osipov K, Valouev IA, Inge-Vechtomov S, Mironova L. N-terminal extension of Saccharomyces cerevisiae translation termination factor eRF3 influences the suppression efficiency of sup35 mutations. FEMS Yeast Res. 2007;7:357–365. doi: 10.1111/j.1567-1364.2006.00176.x. [DOI] [PubMed] [Google Scholar]
 - 43.Hoshino S, Imai M, Kobayashi T, Uchida N, Katada T. The eukaryotic polypeptide chain releasing factor (eRF3/GSPT) carrying the translation termination signal to the 3'-poly(A) tail of mRNA. J. Biol. Chem. 1999;274:16677–16680. doi: 10.1074/jbc.274.24.16677. [DOI] [PubMed] [Google Scholar]
 - 44.Cosson B, Couturier A, Chabelskaya S, Kiktev D, Inge-Vechtomov S, Philippe M, Zhouravleva G. Poly(A)-binding protein acts in translation termination via eukaryotic release factor 3 interaction and does not influence [PSI+] propagation. Mol. Cell. Biol. 2002;22:3301–3315. doi: 10.1128/MCB.22.10.3301-3315.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 45.Volkov KV, Aksenova YA, Soom MJ, Sipov KV, Svitin AV, Kurischko C, Shkundina IS, Ter-Avanesyan MD, Inge-Vechtomov SG, Mironova LN. Novel non-Mendelian determinant involved in the control of translation accuracy in Saccharomyces cerevisiae. Genetics. 2002;160:25–36. doi: 10.1093/genetics/160.1.25. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 46.Jung G, Jones G, Masison DC. Amino acid residue 184 of yeast Hsp104 chaperone is critical for prion-curing by guanidine, prion propagation, and thermotolerance. Proc. Natl. Acad. Sci. USA. 2002;99:9936–9941. doi: 10.1073/pnas.152333299. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 47.Rogozq TM, Viktorovskaia OV, Rodionova SA, Ivanov MS, Volkov KV, Mironova LN. Search for the genes critical for propagation of the prion-like antisuppressor determinant [ISP+] in yeast using insertion library. Mol. Biol. (Mosk) 2009;43:392–399. [PubMed] [Google Scholar]
 - 48.Fingerman I, Nagaraj V, Norris D, Vershon AK. Spf1 plays a key role in yeast ribosome biogenesis. Eukaryotic Cell. 2003;2:1061–1068. doi: 10.1128/EC.2.5.1061-1068.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 49.Caughey B, Race RE, Ernst D, Buchmeier MJ, Chesebro B. Prion protein (PrP) biosynthesis in scrapie-infected and uninfected neuroblastoma cells. J. Virol. 1989;63:175–181. doi: 10.1128/jvi.63.1.175-181.1989. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 50.Deleault NR, Harris BT, Rees JR, Supattapone S. Formation of native prions from minimal components in vitro. Proc Natl Acad Sci U S A. 2007;104:9741–9746. doi: 10.1073/pnas.0702662104. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 51.Bousset L, Beirhali H, Janin J, Melki R, Morera S. Structure of the globular region of the prion protein Ure2 from the yeast Saccharomyces cerevisiae. Structure. 2001;9:39–46. doi: 10.1016/s0969-2126(00)00553-0. [DOI] [PubMed] [Google Scholar]
 - 52.Umland TC, Taylor KL, Rhee S, Wickner RB, Davies DR. The crystal structure of the nitrogen catabolite regulatory fragment of the yeast prion protein Ure2p. Proc. Natl. Acad. Sci. USA. 2001;98:1459–1464. doi: 10.1073/pnas.041607898. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 53.Komar AA, Lesnik T, Cullin C, Merrick WC, Trachsel H, Altmann M. Internal initiation drives the synthesis of Ure2 protein lacking the prion domain and affects [URE3] propagation in yeast cells. EMBO J. 2003;22:1199–1209. doi: 10.1093/emboj/cdg103. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 54.Edskes HK, Gray VT, Wickner RB. The [URE3] prion is an aggregated form of Ure2p that can be cured by overexpression of Ure2p fragments. Proc. Natl. Acad. Sci. USA. 1999;96:1498–1503. doi: 10.1073/pnas.96.4.1498. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 55.Nelson RJ, Ziegilhoffer T, Nicolet C, Werner-Washburne M, Craig EA. The translation machinery and 70 kDal heat shock protein cooperate in protein synthesis. Cell. 1992;71:97–105. doi: 10.1016/0092-8674(92)90269-i. [DOI] [PubMed] [Google Scholar]
 - 56.Chernoff YO, Newnam GP, Kumar J, Allen K, Zink AD. Evidence for a protein mutator in yeast: role of the Hsp70-related chaperone Ssb in formation, stability and toxicity of the [PSI+] prion. Mol. Cell. Biol. 1999;19:8103–8112. doi: 10.1128/mcb.19.12.8103. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 57.Chacinska A, Szczesniak B, Kochneva-Pervukhova NV, Kushnirov VV, Ter-Avanesyan MD, Boguta M. Ssb1 chaperone is a [PSI+] prion-curing factor. Curr Genet. 2001;39:62–67. doi: 10.1007/s002940000180. [DOI] [PubMed] [Google Scholar]
 - 58.Satpute-Krishnan P, Serio TR. Prion protein remodelling confers an immediate phenotypic switch. Nature. 2005;437:262–265. doi: 10.1038/nature03981. [DOI] [PubMed] [Google Scholar]
 - 59.Chattopadhyay S, Das B, Dasgupta C. Reactivation of denatured proteins by 23S ribosomal RNA: role of domain V. Proc. Natl. Acad. Sci. USA. 1996;93:8284–8287. doi: 10.1073/pnas.93.16.8284. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 60.Bach S, Talarek N, Andrieu T, Vierfond JM, Mettey Y, Galons H, Dormont D, Meijer L, Cullin C, Blondel M. Isolation of drugs active against mammalian prions using a yeast-based screening assay. Nat. Biotechnol. 2003;21:1075–1081. doi: 10.1038/nbt855. [DOI] [PubMed] [Google Scholar]
 - 61.Tribouillard-Tanvier D, Beringue V, Desban N, Gug F, Bach S, Voisset C, Galons H, Laude H, Vilette D, Blondel M. Antihypertensive drug guanabenz is active in vivo against both yeast and mammalian prions. PlosOne. 2008;3:e1981. doi: 10.1371/journal.pone.0001981. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 62.Tribouillard-Tanvier D, Dos Reis S, Gug F, Voisset C, Beringue V, Sabate R, Kikovska E, Talarek N, Bach S, Huang C, Desban N, Saupe SJ, Supattapone S, Thuret J-Y, Chedin S, Vilette D, Galons H, Sanyal S, Blondel M. Protein folding activity of ribosomal RNA is a selective target of two unrelated antiprion drugs. Plos One. 2008;3:e2174. doi: 10.1371/journal.pone.0002174. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 63.Sharma D, Masison DC. Hsp70 structure, function, regulation and influence on yeast prions. Prot. & Peptide Lett. 2009;16:571–581. doi: 10.2174/092986609788490230. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 64.Bailey CH, Kandel ER. Synaptic remodeling, synaptic growth and the storage of long-term memory in Aplysia. Prog. Brain Res. 2008;169:179–198. doi: 10.1016/S0079-6123(07)00010-6. [DOI] [PubMed] [Google Scholar]
 - 65.Richter JD. CPEB: a life in translation. TIBS. 2007;32:279–285. doi: 10.1016/j.tibs.2007.04.004. [DOI] [PubMed] [Google Scholar]
 - 66.Si K, Lindquist S, Kandel ER. A neuronal isoform of the Aplysia CPEB has prion-like properties. Cell. 2004;115:879–891. doi: 10.1016/s0092-8674(03)01020-1. [DOI] [PubMed] [Google Scholar]
 - 67.Muhlrad D, Decker CJ, Parker R. Deadenylation of the unstable mRNA encoded by the yeast MFA2 gene leads to decapping followed by 5'->3' digestion of the transcript. Genes Dev. 1994;8:855–866. doi: 10.1101/gad.8.7.855. [DOI] [PubMed] [Google Scholar]
 - 68.Derkatch IL, Bradley ME, Hong JY, Liebman SW. Prions affect the appearance of other prions: the story of [PIN] Cell. 2001;106:171–182. doi: 10.1016/s0092-8674(01)00427-5. [DOI] [PubMed] [Google Scholar]
 - 69.Du Z, Park K-W, Yu H, Fan Q, Li L. Newly identified prion linked to the chromatin-remodeling factor Swi1 in Saccharomyces cerevisiae. Nat. Genet. 2008;40:460–465. doi: 10.1038/ng.112. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 70.Nemecek J, Nakayashiki T, Wickner RB. A prion of yeast metacaspase homolog (Mca1p) detected by a genetic screen. Proc. Natl. Acad. Sci. USA. 2008;106:1892–1896. doi: 10.1073/pnas.0812470106. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
 - 71.Patel BK, Gavin-Smyth J, Liebman SW. The yeast global transcriptional co-repressor protein Cyc8 can propagate as a prion. Nat. Cell Biol. 2009;11:344–349. doi: 10.1038/ncb1843. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 72.Coustou V, Deleu C, Saupe S, Begueret J. The protein product of the het-s heterokaryon incompatibility gene of the fungus Podospora anserina behaves as a prion analog. Proc. Natl. Acad. Sci. USA. 1997;94:9773–9778. doi: 10.1073/pnas.94.18.9773. [DOI] [PMC free article] [PubMed] [Google Scholar]
 - 73.Alberti S, Halfmann R, King O, Kapila A, Lindquist S. A systematic survey identifies prions and illuminates sequence features of prionogenic proteins. Cell. 2009;137:146–158. doi: 10.1016/j.cell.2009.02.044. [DOI] [PMC free article] [PubMed] [Google Scholar]
 


