Skip to main content
  • Protein family review
  • Published:

Poly(A)-binding proteins: multifunctional scaffolds for the post-transcriptional control of gene expression

Summary

Most eukaryotic mRNAs are subject to considerable post-transcriptional modification, including capping, splicing, and polyadenylation. The process of polyadenylation adds a 3' poly(A) tail and provides the mRNA with a binding site for a major class of regulatory factors, the poly(A)-binding proteins (PABPs). These highly conserved polypeptides are found only in eukaryotes; single-celled eukaryotes each have a single PABP, whereas humans have five and Arabidopis has eight. They typically bind poly(A) using one or more RNA-recognition motifs, globular domains common to numerous other eukaryotic RNA-binding proteins. Although they lack catalytic activity, PABPs have several roles in mediating gene expression. Nuclear PABPs are necessary for the synthesis of the poly(A) tail, regulating its ultimate length and stimulating maturation of the mRNA. Association with PABP is also a requirement for some mRNAs to be exported from the nucleus. In the cytoplasm, PABPs facilitate the formation of the 'closed loop' structure of the messenger ribonucleoprotein particle that is crucial for additional PABP activities that promote translation initiation and termination, recycling of ribosomes, and stability of the mRNA. Collectively, these sequential nuclear and cytoplasmic contributions comprise a cycle in which PABPs and the poly(A) tail first create and then eliminate a network of cis- acting interactions that control mRNA function.

Gene organization and evolutionary history

RNA-binding proteins are often purified and classified on the basis of the RNA sequences with which they interact [1]. One class of these factors comprises proteins recognizing the homopolymeric polyadenylate tracts that are added to the 3' end of most mRNAs. Poly(A)-binding proteins have been identified in many eukaryotes, but appear to be absent from prokaryotes. PABP genes have been cloned from a number of organisms, and their sequences are available in several databases; a current list with database links is available as an additional data file and on our website [2]. Typically, only one gene encoding cytoplasmic PABP (PABPC) is present in the single-cell eukaryotes, whereas multiple PAPBC genes are present in metazoans and plants (Table 1, Figure 1). A single gene encoding a nuclear PABP (PABPN) has also been identified in cow, frog, human, mouse, fly, worm, and yeasts (Figure 1). A phylogenetic analysis comparing all known PABP protein sequences groups PABPs by organism type (such as metazoans, yeast, and plants) and also identifies similarities among the PABP family members (Figure 1). To date, genes encoding a single nuclear PABP and four cytoplasmic PABPs, as well as four pseudogenes, have been identified in human cells, and their chromosomal locations have been mapped (Table 2). In humans, three lineages of PABP proteins are observed: cytoplasmic PABPs (PABPC1, PABPC3, and iPABP); nuclear PABP (PABPN1); and X-linked PABP (PABPC5). Within the PABPC group, PABPC1 and PABPC3 are most closely related. Interestingly, the mouse gene encoding the alternate PABP, mPABPC2, seems to be a retroposon, as it has no introns and its promoter is distinct from that of mPABPC1 [3]; mPABPC2 is most closely related to hPABPC3, which also lacks introns [4]. Similarly, all the characterized PABPC5 genes lack introns [5], suggesting that they too may be derived from retrotransposition events.

Figure 1
figure 1

Predicted evolutionary relationships of PABPs. Full-length PABP sequences were compiled from various databases (see Additional data files) and aligned using the CLUSTALW program at the European Bioinformatics Institute [122]. The tree was constructed using the neighbor-joining method [123] and drawn using Phylodendron [124]. The scale bar represents 0.1 substitutions. In the instances where no PABP name is given, only a single PABP protein has been identified in that organism.

Table 1 Genes encoding cytoplasmic PABPs in various organisms
Table 2 Chromosomal location of human PABP genes

A comparable evolutionary analysis was reported for the eight PAB genes identified in the plant Arabidopsis thaliana [6]. Phylogenetic comparisons coupled with expression analyses identified four classes of PABP proteins. In class I (PAB3 and PAB5), expression is limited to reproductive tissue; class II members (PAB2, PAB4 and PAB8) are highly and broadly expressed; class III PABPs (PAB6 and PAB7) have a restricted, weak expression pattern; and the sole member of class IV (PAB1) has low, tissue-specific expression. Comparison of the Arabidopsis PABPs with those from rice indicates that the duplication events which gave rise to classes I-III in flowering plants occurred prior to the divergence of monocots and dicots, more than 200 million years ago [6]. By analyzing the conservation and loss of introns within the PABP gene family, an evolutionary model has been derived in which an ancestral PABP independently gave rise to classes II, III and IV, with class I subsequently derived from class II [6]. Although all eight of the Arabidopsis PABPs are more closely related to the set of nuclear PABPs than to the PABPs of most other eukaryotes (Figure 1), none of these proteins appears to be an authentic PABPN1 species.

One interesting characteristic conserved among the PABPC1 genes is an adenylate-rich region in the 5' untranslated region (UTR). Several studies have suggested that PABP regulates its own expression by binding to these sequences [7–9].

Characteristic structural features

The association of PABPs with poly(A) requires a minimal binding site of 12 adenosines, and multiple PABP molecules can bind to the same poly(A) tract, forming a repeating unit covering approximately 27 nucleotides [10–13]. In vitro binding affinities of PABP for poly(A) are of the order of 2-7 nM [13–15]. PABPs interact with poly(A) via RNA-recognition motifs (RRMs; Figure 2).

Figure 2
figure 2

The domains of human PABPs. PABPC1, PABPC3, iPABP, PABPC5 and PABPN1 are shown, aligned on their first RNA-recognition motifs (RRMs). White capsules represent individual RRMs; black hexagons (5H) represent the five conserved helices at the carboxyl terminus. Inverted brackets indicate the site of expansion of a run of alanines in PABN1 that leads to the synthesis of PABPN1 with 12-17 alanines and results in the autosomal recessive disease oculopharyngeal muscular dystrophy (OMPD) [24, 25]. PABPN1 accumulates in OMPD patients and forms intranuclear inclusions that appear to sequester mRNAs and associated factors and promote cell death [25].

The RRM is the most prevalent domain used in the recognition of RNA, as shown by its presence in hundreds of different proteins [16]. RRMs, which are typically 90-100 amino acids in length, appear to be present in proteins in all types of organisms, suggesting that this is an ancestral motif with important functions in RNA biology. Solution nuclear magnetic resonance (NMR) and X-ray crystallographic studies have determined that the RRM is a globular domain composed of a four-stranded anti-parallel β sheet backed by two α helices (Figure 3a) [17]. The central two β strands of each RRM include two highly conserved sequence motifs, octameric RNP1 ((K/R)-G-(F/Y)-(G/A)-F-V-X-(F/Y), where X is any amino acid) and hexameric RNP2 ((L/I)-(F/Y)-(V/I)-(G/K)-(N/G)-(L/M)) (Figure 3a). The electron density map of the human PABPC1-oligo(A) complex identifies eight adenylate residues extending through a trough lined by the β-sheets of the RNPs (Figure 3b) [17]. Specificity for recognition of poly(A) is primarily mediated via van der Waals contacts, hydrogen bonds, and stacking interactions with conserved residues within the RNP motifs [17].

Figure 3
figure 3

Structures of the domains of human PABPC1. (a) Crystal structure of RRMs 1 and 2 in association with poly(A) [17]. The central two 3 strands of each RRM include two highly conserved sequence motifs, octomeric RNP1 ((K/R)-G-(F/Y)-(G/A)-F-V-X-(F/Y), where X is any amino acid) and hexameric RNP2 ((L/I)-(F/Y)-(V/I)-(G/K)-(N/G)-(L/M)), which is repeated six times. (b) The RNA-binding trough that is present when RRM1 and RRM2 of human associate with poly(A). (c) NMR structure of the five carboxy-terminal helices [21]. Figures were generated by MOLSCRIPT 2.0 using data from Protein data bank (PDB) files (a,b) 1CVJ and (c) 1G9L [125, 126].

Cytoplasmic PABPs

The overall structure of the cytoplasmic PABPs is highly conserved and consists of four RRMs connected to a carboxy-terminal helical domain by an unstructured linker region rich in proline and methionine residues [12, 18]. Phylogenetic analysis suggests that the four RRMs arose from successive duplications before the divergence of yeast and mammals [19]. The first two RRMs make up one functional unit and the latter two make up a second. This conclusion is derived partly from the observation that residues participating in RNA recognition within RRM1 are most similar to those in RRM3, while those of RRM2 are most like those of RRM4 [17]. Although each RRM is capable of binding RNA, they are not functionally equivalent, as they have differing affinities for poly(A) [15].

The carboxy-terminal helical domain is highly conserved. In humans it is composed of five helices (Figure 3c), while the yeast protein has only four, lacking an ortholog of the first helix [20, 21]. The carboxy-terminal domain is not required for RNA recognition, is dispensable for cell viability in yeast [13, 15], and is missing from PABPC5 proteins [5]. This domain is shared with HECT domain proteins in the hyperplastic disc (HYD) family of ubiquitin-protein ligases [22], but there is no evidence that PABPs play any role in protein degradation. The carboxy-terminal domain is, however, the site of interaction with factors regulating polyadenylation, deadenylation, translation initiation, and translation termination (see below).

Nuclear PABPs

The structure of the nuclear PABPs is not as well understood as that of the cytoplasmic PABPs, largely because crystal and NMR structures have yet to be determined, but it is known that they typically have an acidic amino terminus followed by a single RRM and an arginine-rich carboxy-terminal domain. Recognition of poly(A) requires both the RRM and the arginine-rich domain [23]. A run of alanines in PABN1 is expanded in the recessive disease oculopharyngeal muscular dystrophy (see Figure 2) [24, 25].

In yeast, the nuclear PABP is essential for viability and is encoded by the NAB2 gene [26]. Unlike other poly(A)-binding proteins, Nab2p uses an Arg-Gly-Gly (RGG) domain for binding. This protein also contains a Cys-Cys-Cys-His zinc-binding motif, similar to one in RNA polymerase subunits, and a glutamine-rich region that contains a variable number of Gln-Gln-Gln-Pro segments, the number of which is strain-dependent.

Localization and function

PABPs have crucial roles in the pathways of gene expression. They bind the poly(A) tails of newly synthesized or mature mRNAs and appear to act as cis-acting effectors of specific steps in the polyadenylation, export, translation, and turnover of the transcripts to which they are bound. Lacking any evident catalytic activity, PABPs provide a scaffold for the binding of factors that mediate these steps and also apparently act as antagonists to the binding of factors that enable the terminal steps of mRNA degradation.

Polyadenylation

Messenger RNAs synthesized in the nucleus generally contain a 3' poly(A) tail; the rare exceptions to this rule are principally the transcripts of replication-dependent histone genes. Newly synthesized poly(A) tails of different mRNAs are relatively homogeneous in length and approximately 200-250 residues in mammals and 70-90 residues in yeast [27]. These poly(A) tracts are not encoded within genes but are added to nascent pre-mRNAs in a two-step processing reaction that involves site-specific cleavage and subsequent polyadenylation of the upstream cleavage product [23, 28–30]. Throughout eukaryotes, pre-mRNA cleavage and polyadenylation take place in a large complex (500-1,000 kDa) that includes poly(A) polymerase (PAP) and many additional factors. In general, the factors regulating PAP stimulate both the specificity and processivity of an otherwise marginally active and indiscriminate enzyme. In so doing, they not only regulate the process of polyadenylation but also determine the ultimate size of the poly(A) tail.

In mammalian cells, PABPN1 binds nascent tracts of 11-14 adenylate residues [31] and, along with cleavage and polyadenylation specificity factor (CPSF), stimulates PAP to switch from distributive synthesis (dropping off after synthesis of a few nucleotides) to processive (continuous, highspeed) synthesis [32, 33]. PABPN1 monomers continue to bind available, nascent adenylates until the full-length poly(A) tail has been synthesized and the polymerase then reverts back to its distributive mode [34]. This sequential binding is accompanied by the formation of linear filaments and 21 nm spherical particles: the latter are thought to serve as 'molecular rulers' that dictate the final length of the poly(A) tail [34]. In this model, the particle is postulated to encompass a stable polyadenylation complex and to tolerate PABPN1-poly(A) oligomers until the tail reaches 200-300 nucleotides. Beyond that point, increased poly(A) length is believed to be compromised by disruption of critical interactions between PAP and CPSF [34].

PABPs also play a role in the polyadenylation of yeast pre-mRNAs. Recent studies indicate that Nab2p is the most likely candidate for the yeast equivalent of PABPN1 function, at least for a subset of mRNAs. Mutations in NAB2 promote hyperpolyadenylation of mRNA that cannot be reversed by overexpression of Pab1p [35]. The failure to detect this activity of Nab2p in earlier studies may be attributable to inhibitory interactions between Nab2p and its nuclear import receptor Kap104p, and/or to the preponderance of Pab1p in whole-cell extracts used for in vitro polyadenylation and the consequent obstruction of Nab2p activity by Pab1p bound to nascent poly(A) [35]. Interestingly, mutations in the yeast gene encoding cytoplasmic PABP, PAB1, cause a significant increase in mRNA poly(A) tail lengths in vivo and in vitro [36–38], and this effect, too, is partly attributable to a switch of PAP (Pap1p) between processive and distributive activities. Unlike the process in mammalian cells, the yeast switch appears to be directly regulated by Fip1p and Yth1p, two factors unrelated to nuclear or cytoplasmic PABPs, and only indirectly regulated by Pab1p [39–41]. Pab1p interactions underlying this indirect effect may include its binding to the nascent mRNA [28] or a direct interaction with the RNA-processing factor Rna15p [37].

Evidence for a direct role for Pab1p in yeast poly(A) length control comes from experiments analyzing the Pab1p-mediated regulation of poly(A) nuclease (PAN). This exonuclease, comprising the Pan2p and Pan3p proteins, appears to trim up to 20 residues from excessively long newly synthesized poly(A) tails in an mRNA-specific manner [42–44]. Pan2p, the subunit with apparent exonuclease activity, is positively and negatively regulated by interactions with Pan3p and Pbp1p, respectively; both of the latter interact with Pab1p (D.M. and A.J., unpublished observations; [43–45]).

Nuclear export

A second role for PABPs in the nuclear maturation of mRNA can be inferred from experiments in which impaired 3' processing interferes with export of mRNAs to the cytoplasm. In both mammalian cells and yeast, mRNAs are generally retained in the nucleus when they lack a functional polyadenylation signal or when polyadenylation is inhibited by the absence or inactivity of specific catalytic factors [46–49]. Since the failure to polyadenylate an mRNA would deprive it of bound PABPs, nuclear retention of mRNA could be attributable to an essential role for PABPs in mRNA export.

As noted above, PABPs coat the nascent poly(A) tail and play a role in determining its ultimate length. How, then, might this poly(A)-PABP complex facilitate the exit of mRNAs and their associated proteins (mRNPs) from the nucleus? Consistent with the propensity of PABPs to form interactions critical to specific functions, yeast Pab1p has been shown to interact with specific nucleoporins [50] and the nuclear export signal export receptor, Xpo1p [49], and Nab2p has been shown to interact with Gfd1p, a nuclear-pore-associated protein [51]. The presence of bound Pab1p or Nab2p could serve as a determinant of an mRNP's export competence, in a manner analogous to the function of the RNA export factor Yra1p [52]. This view is consistent with the observed nucleocytoplasmic shuttling of yeast and mammalian PABPs [53–56] and with the inhibitory effects on mRNA export caused by interactions between the influenza virus NS1A protein and PABPN1 [57].

The notion of a direct role for PABPs in mRNA export may, however, be too simplistic. It does not accommodate examples of mRNAs that enter the cytoplasm without conventional 3' processing [58, 59], viable mutants devoid of PABP [36], or functional interactions between the 3' processing apparatus and the factors that promote mRNA export [49, 60]. The latter reflect a quality control mechanism that leads to retention of an mRNA in the nucleus (often at its transcription site) in the event of processing problems [49, 61, 62]. This apparent checkpoint illustrates the interdependence of many steps in gene expression and the manner in which such regulatory mechanisms can make indirect effects appear to be direct.

Translation initiation

After an mRNA enters the cytoplasm, the association of PABP with its poly(A) tail promotes 5'-3' interactions that stimulate initiation of its translation [27, 63]. Formation of this 'closed loop' [27] was shown by Sachs and colleagues [64–66] to promote the recruitment of 40S ribosomal subunits and to be dependent, at a minimum, on interactions between initiation factor eIF4G and PABP and concurrent interactions between eIF4G and the cap-binding protein eIF4E (Figure 4). The existence of a translational regulatory network involving PABP, eIF4G, and eIF4E is consistent with the impaired-translation phenotypes of yeast strains lacking functional Pab1p [36] and provides a mechanistic basis for the synergistic effects on translation known to occur when mRNAs are both capped and polyadenylated [65, 67, 68]. The combined cooperative interactions enhance the affinity of eIF4E for the 5' cap of the mRNA by lowering its dissociation rate [69–72], stimulate the RNA-binding activity of PABP [73], and increase the ATPase and RNA helicase activities of eIF4A, eIF4B, and eIF(iso)4F [74]. The combination of these effects also provides an effective means for the protein synthesis apparatus to ensure preferential translation of mRNAs containing both a cap and a poly(A) tail [74] and may create an opportunity for ribosomes to recycle from the 3' to the 5' end of the same mRNA [27, 75].

Figure 4
figure 4

Roles of PABP in mRNA translation and stability. This model depicts different stages of a cytoplasmic mRNA 'life cycle', in which distinct roles can be ascribed to PABP. (a) Association of PABP with the mRNA poly(A) tail. (b) Interaction of PABP with elongation initiation factor eIF4G to promote formation of the 'closed loop', thus (c) initiating translation and antagonizing decapping. (d) Interaction of PABP with the termination factor eRF3 and recycling of the ribosome from the 5' to the 3' end of the same mRNA. (e) Poly(A) shortening by the Ccr4p-Pop2p-Notp deadenylase complex. (f) Loss of the poly(A) tail and PABP, facilitating (g) dissociation of the proteins of the mRNP, binding of the Lsm1-7p-Pat1p complex, and decapping by the decapping proteins Dcp1p and Dcp2p, and subsequent (h) 5'-to-3' degradation of the mRNA by the exonuclease Xrn1p or (i) 3'-to-5' degradation by the exosome.

Studies in yeast and mammalian cells have shown that the Pab1p-eIF4G interaction requires RRM1 and RRM2 of Pab1p (the same RRMs required for poly(A) recognition) and an amino-terminal domain of eIF4G [65, 76–78]. Several additional experiments have indicated, however, that the network of 5'-3' interactions regulating translation initiation goes well beyond the communication of a single domain in PABP with another in eIF4G. This was initially suggested by the existence of viable yeast pab1 mutants in which the Pab1p-eIF4G interaction could not occur [65] and others that had defects in poly(A)-dependent translation but no defects in eIF4G binding [76]. The potential complexity of PABP's translation-promoting interactions is illustrated by interactions of PABPs in wheat germ with the initiation factor eIF4B [73] and in mammals with the PABP-interacting proteins Paip1 and Paip2 [79–82]. Paip1 is homologous to the central segment of mammalian eIF4G and binds with high affinity and 1:1 stoichiometry to two sites in PABP, one in RRMs 1 and 2 and the other in the carboxy-terminal domain [79, 80]. The region of eIF4G to which Paip1 is homologous encompasses one of two binding sites for the RNA helicase eIF4A. Not surprisingly, Paip1 also interacts with eIF4A, and is capable of stimulating the translation of a reporter mRNA when overexpressed in cultured cells [79]. Paip2, a low-molecular-weight acidic protein, binds PABP at two sites, one in RRMs 2 and 3 and one in the carboxyl terminus [81, 82]. Binding of Paip2 to the RRM2-3 region competes effectively for binding of Paip1 to PABP, reduces PABP binding to poly(A), and inhibits the translation of polyadenylated mRNA [81, 82].

Tethered-function assays in yeast and Xenopus that exploit PABP fusions to the bacteriophage MS2 coat protein also underscore the intricate nature of PABP's stimulatory effects on translation [83]. PABP tethered at specific MS2 coat binding sites stimulates translation of a reporter mRNA in cis, but not in trans, and can do so without its poly(A)-binding activity and in the absence of a poly(A) tail [83]. With the exception of the yeast requirement that Pab1p be bound to poly(A) in order to interact with eIF4G [64], this implies that, at least with respect to translational stimulation, poly(A) simply provides a binding site for PABP. The failure of yeast Pab1p to function in the absence of bound poly(A) may reflect the selective inability of yeast eIF4G to stabilize the packing of poly(A)-associated RRMs 1 and 2 in a manner comparable to that achieved by the eIF4Gs of other species [84]. Tethered function assays also reveal that RRMs 1 and 2, or RRMs 3 and 4, of Xenopus PABP are as capable of translational stimulation as the full-length protein, despite the fact that RRMs 3 and 4 lack the ability to interact with eIF4G or Paip1 [83]. Like the Pab1p-eIF4G interaction mutants in yeast [65, 76], the novel PABP interactors in mammals and plants [73, 79–82], and the unique domain requirements for transactivation of translation by Pab1p [78], this observation implies that interaction with eIF4G is not likely to be the only mechanism by which PABP stimulates translation. One alternative model for PABP function, supported by genetic analyses in yeast [85] and the biochemical properties of poly(A)-deficient mRNAs in vitro [67], suggests that PABP is also a regulator of the joining of the 60S subunit to the 40S preinitiation complex. The studies in yeast indicate that PABP controls 60S joining by regulating the activities of two RNA heli-cases, Ski2p and Slh1p [85].

Additional insight into the translational networks affected by the presence of PABP is derived from studies of the tactics that viruses and cells use to modulate PABP structure and/or activity. For example, rotaviruses reroute translation for their own purposes by synthesizing a protein, NSP3, which serves as a PABP analog. NSP3 binds to specific 3' sequences on viral mRNAs and effectively circularizes those transcripts, and mimics PABP, by also binding to eIF4G [86]. Enteroviruses, on the other hand, choose to eliminate the activity of PABP, rather than replace it. As part of a general assault on host cap-dependent translation, these viruses express two proteases, 2A and 3C, that not only remove the PABP-interacting domain of eIF4G but also cleave PABP into several fragments [87, 88]. PABP interactions and activity, at least in plants, are also altered by changes in its phosphorylation status [89] and may be affected by arginine methylation within the domain separating RRM4 from the carboxy-terminal helices [90].

In addition to their global effects on translation initiation, PABPs can also selectively affect the translation of individual mRNAs. PABPs can bind oligoadenylate tracts in the 5' UTRs of their own mRNAs, thereby repressing their own translation (and possibly their stability [17]) [9, 91]. This autoregulation can be mimicked both in vitro and in vivo, can be abolished by deletion of the adenylate-rich region and can be conferred on other mRNAs by insertion of the adenylate-rich tract within their 5' UTRs. In each case, the presence of PABP is required to mediate the observed effects. The inhibition of translation has been ascribed to an inability of the 60S ribosomal subunit to join the pre-initiation complex [92]. PABP can also facilitate the binding of translational repressors specific for other mRNAs, such as that encoding the iron-oxidizing protein ceruloplasmin [93], and can activate the translation of a large number of mRNAs whose polyadenylation is developmentally controlled [94], as well as functioning as an mRNA-specific activator. Cytoplasmic PABP in Chlamydomonas reinhardtii, normally a 69 kDa polypeptide, is imported into chloroplasts where it is processed to a 47 kDa form that binds the 5' UTR of the psbA mRNA and activates its translation [95]. The latter role of PABP is particularly intriguing in light of the generally prokaryotic nature of chloroplast translation systems.

Translation termination

The eukaryotic translation termination factor eRF1, which is responsible for catalyzing polypeptide hydrolysis in response to recognition of any of the three nonsense codons by the ribosome, appears to be activated by the GTPase eRF3 [96]. The amino-terminal region of eRF3 does not participate in this interaction with eRF1, but does interact directly with the carboxy-terminal domain of cytoplasmic PABPs [21, 97, 98]. The eRF3-PABP interaction appears to enhance the efficiency of termination in cells with mutated or aggregated eRF3 [98] and to promote ribosome recycling for multiple rounds of translation on the same mRNA [99]. It also seems to minimize the multimerization of PABP monomers on poly(A), possibly expediting access of poly(A) shortening enzymes to their substrate and linking translational termination to normal mRNA decay [97]. Additional insights into the role of PABPs in translation termination come from analyses of instances in which termination occurs abnormally, such as at premature nonsense codons. In this case, termination is thought to be aberrant because of the creation of a 'faux' UTR, an untranslated region lacking at least one of the factors required for efficient polypeptide hydrolysis and ribosome release that are normally positioned 3' to a termination codon by interaction with poly(A)-associated PABP [100].

Decay of mRNA

The process of mRNA decay can be initiated by three distinct events: endonucleolytic cleavage, removal of the 5' cap, and poly(A) shortening [101]. In yeast, in which the process of mRNA decay has been extensively analyzed, most wild-type mRNAs decay by a mechanism in which the initial nucleolytic event is the shortening of the poly(A) tail to an oligo(A) length of 10-15 nucleotides. After poly(A) shortening, transcripts are decapped by the Dcp1p-Dcp2p complex. Decapped and deadenylated mRNAs are then digested exonucleolytically by the 5'-to-3' exoribonuclease, Xrn1p, and/or the 3'-to-5' multi-subunit exosome [102] (Figure 4).

All three decay-initiating events eliminate the closed-loop state of the mRNP by removing or separating the binding sites for the respective 5' and 3' interacting proteins [27], and these events also render the remaining mRNA fragments substrates for further degradation. At a minimum, then, mRNA decay generally occurs concurrently with the conversion of an mRNP from a translatable to an untranslatable (or poorly translatable) form [27, 101], that is, in parallel with the termination of PABP's role in the enhancement of translation initiation.

Although the onset of mRNA decapping does coincide with the loss of PABP's binding site, and efficient translation initiation does, indeed, antagonize mRNA decay [103, 104], PABP's role in the maintenance of mRNA stability is more complicated than that of a mere translation enhancer. Several observations suggest that loss of PABP's binding site, and presumptive disruption of the closed loop state, may not always trigger immediate degradation of the remainder of the mRNA. These observations include, first, that poly(A) shortening or removal is the rate-determining event in the decay of some mRNAs whereas for others, it may be an obligate event in their degradation but not the rate-determining step [105, 106]; second, that yeast pab1 mutations that unlink mRNA decapping from poly(A) shortening do not necessarily accelerate the rate of mRNA decay [107, 108]; and third, that the domains of tethered Pab1p that provide yeast mRNA stabilization and translation functions are different [83, 109].

Additional roles for PABP in the regulation of mRNA stability range from being an antagonist or promoter of poly(A) shortening to a facilitator of the binding of additional factors that promote or retard rapid mRNA decay. In vitro, excess poly(A) is an effective competitor of PABP binding to mRNA [63, 110–112]. Such competition accelerates the rate of poly(A) shortening, indicating that the presence of PABP on the poly(A) tail provides a protective effect [110–112]. This effect is, in part, attributable to physical hindrance of the deadenylase, because poly(A) tails are often shortened in discrete lengths equivalent in size to a PABP 'footprint' [113]. It is also known, however, that the principal yeast deadenylase, the Ccr4p-Pop2p-Notp complex (Figure 4), and a major mammalian deadenylase, PARN, are both inhibited in the presence of PABP [114]. In contrast, the yeast Pan2p-Pan3p deadenylase, an enzyme responsible for the initial trimming of the poly(A) tail (see above), requires Pab1p for its activity [42, 44].

Much like its role in translational initiation, PABP also influences mRNA decay by interacting with key regulatory proteins, either influencing their activity or being influenced by them. Two proteins that bind the 3' UTR of the α-globin mRNA and enhance its stability, αCP1 and αCP2, interact with human PABP [113]. PABP appears to stimulate the ability of the αCP proteins to bind to their target sequence in the 3' UTR, thereby precluding access of an endonuclease (ErEN) to its recognition site [115]. Interestingly, the binding of PABP to the poly(A) tail is also enhanced by the αCP proteins, implying that α-globin mRNA stabilization is mediated by multiple interdependent events [115]. Stability of the mRNA encoding the transcription factor c-Fos is regulated by sequence elements in its 3' UTR and coding region [116]. The coding region stability element, also known as the major protein-coding-region determinant (mCRD), interacts with a complex of RNA-binding proteins that includes PABP, Paip1, hnRNPD, NSAP1, and Unr [117]. Translation through the mCRD destabilizes c-fos mRNA by a mechanism that is thought to disrupt interactions with this complex and, in turn, promote poly(A) shortening [117]. As noted above, PABP also interacts with the termination factor eRF3 [21, 97, 98], a consequence of which is a decrease in the number of PABP multimers associated with the poly(A) tail. This observation links translation termination to poly(A) shortening and suggests one mechanism for orchestrating a standardized 'clock' that limits the lifetime of a poly(A) tail and, in turn, the mRNA to which it is appended [97].

Additional roles for PABP in mRNA decay are illustrated by events that occur after the poly(A) tail has been removed. As shown in Figure 4, mRNA deadenylation is accompanied by an mRNP rearrangement that allows binding of a decapping activator complex containing the proteins Lsm1p-Lsm7p and Pat1p [118]. This complex appears to promote interaction of the mRNP with the Dcp1p-Dcp2p decapping complex, thereby creating a substrate for terminal 5'-to-3', and/or 3'-to-s' exonucleolytic degradation [119] (Figure 4). Recent studies indicate that all steps subsequent to association of the Lsm1-7p-Pat1p complex occur at a limited number of subcellular sites called P bodies [119]. In principle, therefore, both the terminal steps of mRNA decay (from decapping onwards), and their localization to a specific subcellular site, are prevented from occurring by the presence of bound Pab1p. Pab1p may simply maintain the mRNP in its translation-favorable mode, but the formal possibility that it directly inhibits mRNA association of the Lsm1-7p-Pat1p complex has not been excluded. In the latter hypothesis, PABP's exit from the mRNP would complete a cycle in which its initial association with mRNA assists in mRNP formation, then leads to efficient mRNP utilization, and culminates in destruction of the mRNA.

Frontiers

Considering the number of functions associated with the PABPs, and their simultaneous interactions with both RNA and other proteins, the number of questions for which we have no answers far exceeds the number of those for which we do. Does the presence or absence of PABP determine an mRNP's competence for export, or does it play a more active role? How is nuclear PABP exchanged for cytoplasmic PABP and where does that exchange take place? How does interaction with PABP actually affect eIF4G and eRF3, and vice versa; in other words, do these proteins influence each other's conformations and interactions with other factors? Does autoregulatory PABP simply 'block' the 5' UTR or does it promote interactions with other factors that are the ultimate regulators? Why do plants have so many PABP genes? Have they separated PABP's functions into distinct polypep-tides?

A key question is which of PABP's many functions are essential. Of the functions enumerated in this review, most appear to be dispensable. For example, PABP mutants lacking the ability to interact with factors governing polyadenylation, translation initiation, and translation termination are all viable. Cross-species complementation experiments assessing the essential nature of PABP demonstrate that Arabidopsis Pab3p can restore PABP's role in mRNA biogenesis but fails to complement defects in mRNA decay and translation initiation [120]. What does appear to be required is the ability of PABP to recognize RNA. The possibility remains that the essential nature of PABP lies not with a single function but with a combination of functions. That, of course, raises the final question: have all of PABP's functions been enumerated? That seems unlikely. A hint of PABP's untapped versatility is apparent from its role in the replication of zucchini yellow mosaic potyvirus, a plant virus whose RNA-dependent RNA polymerase appears to exploit PABP for viral replication [121]. Who knows - maybe PABP will find its way into splicing and transcription, completing its act as the one-man band of gene expression.

Additional data files

A list of the currently known PABP genes with accession numbers and links to their entries in the nucleotide and protein sequence databases (Additional data file 1) and the sequences of these proteins in FASTA format (Additional data file 2) are available with the online version of this article.

References

  1. Burd CG, Matunis EL, Dreyfuss G: The multiple RNA-binding domains of the mRNA poly(A)-binding protein have different RNA-binding activities. Mol Cell Biol. 1991, 11: 3419-3424. Characterization of the RNA-binding domains of the yeast poly(A)-binding protein.

    PubMed  PubMed Central  Google Scholar 

  2. The Jacobson lab - additional data. Fasta data and links to poly(A)-binding protein sequences are available on this website., [http://jacobsonlab.umassmed.edu/cgi-bin/dbcontents.cgi?gencon-tents=PABPs]

  3. Kleene KC, Mulligan E, Steiger D, Donohue K, Mastrangelo MA: The mouse gene encoding the testis-specific isoform of poly(A) binding protein (Pabp2) is an expressed retroposon: intimations that gene expression in spermatogenic cells facilitates the creation of new genes. J Mol Evol. 1998, 47: 275-281. This article reports the characterization of mouse testis-specific PABP2 isoform.

    PubMed  Google Scholar 

  4. Feral C, Guellaen G, Pawlak A: Human testis expresses a specific poly(A)-binding protein. Nucleic Acids Res. 2001, 29: 1872-1883. 10.1093/nar/29.9.1872. This article describes the characterization of human testis-specific PABP3 isoform.

    PubMed  PubMed Central  Google Scholar 

  5. Blanco P, Sargent CA, Boucher CA, Howell G, Ross M, Affara NA: A novel poly(A)-binding protein gene (PABPC5) maps to an X-specific subinterval in the Xq21.3/Yp11.2 homology block of the human sex chromosomes. Genomics. 2001, 74: 1-11. 10.1006/geno.2001.6530. The human PABPC5 was localized to the X chromosome.

    PubMed  Google Scholar 

  6. Belostotsky DA: Unexpected complexity of poly(A)-binding protein gene families in flowering plants: three conserved lineages that are at least 200 million years old and possible auto- and cross-regulation. Genetics. 2003, 163: 311-319. An article describing three distinct classes of PABP genes in Arabidopsis.

    PubMed  PubMed Central  Google Scholar 

  7. de Melo Neto OP, Standart N, Martins de Sa C: Autoregulation of poly(A)-binding protein synthesis in vitro. Nucleic Acids Res. 1995, 23: 2198-2205. Binding of PABP to an A-rich tract in the 5' UTR regulates PABP gene expression.

    PubMed  PubMed Central  Google Scholar 

  8. Bag J, Wu J: Translational control of poly(A)-binding protein expression. Eur J Biochem. 1996, 237: 143-152. PABP binding to the 5' UTR of its own mRNA to regulate translation in cis.

    PubMed  Google Scholar 

  9. Wu J, Bag J: Negative control of the poly(A)-binding protein mRNA translation is mediated by the adenine-rich region of its 5'-untranslated region. J Biol Chem. 1998, 273: 34535-34542. 10.1074/jbc.273.51.34535. The authors report the autoregulation of PABP expression in HeLa cells.

    PubMed  Google Scholar 

  10. Baer BW, Kornberg RD: Repeating structure of cytoplasmic poly(A)-ribonucleoprotein. Proc Natl Acad Sci USA. 1980, 77: 1890-1892. Demonstration of the 27-nucleotide repeats in PABP/poly(A) complexes.

    PubMed  PubMed Central  Google Scholar 

  11. Baer BW, Kornberg RD: The protein responsible for the repeating structure of cytoplasmic poly(A)-ribonucleoprotein. J Cell Biol. 1983, 96: 717-721. Demonstration that 27-nucleotide RNP repeats are attributable to PABP.

    PubMed  Google Scholar 

  12. Sachs AB, Bond MW, Kornberg RD: A single gene from yeast for both nuclear and cytoplasmic polyadenylate-binding proteins: domain structure and expression. Cell. 1986, 45: 827-835. The cloning and characterization of the yeast PAB1 gene.

    PubMed  Google Scholar 

  13. Sachs AB, Davis RW, Kornberg RD: A single domain of yeast poly(A)-binding protein is necessary and sufficient for RNA binding and cell viability. Mol Cell Biol. 1987, 7: 3268-3276. An analysis of the consequences of large deletions in the yeast PAB1 gene.

    PubMed  PubMed Central  Google Scholar 

  14. Gorlach M, Burd CG, Dreyfuss G: The mRNA poly(A)-binding protein: localization, abundance, and RNA-binding specificity. Exp Cell Res. 1994, 211: 400-407. 10.1006/excr.1994.1104. Specific monoclonal antibodies were used to determine the subcellular localization, concentration, and binding affinities of human PABP.

    PubMed  Google Scholar 

  15. Kuhn U, Pieler T: Xenopus poly(A)-binding protein: functional domains in RNA binding and protein-protein interaction. J Mol Biol. 1996, 256: 20-30. 10.1006/jmbi.1996.0065. Functional characterization of distinct domains within Xenopus PABP.

    PubMed  Google Scholar 

  16. Burd CG, Dreyfuss G: Conserved structures and diversity of functions of RNA-binding proteins. Science. 1994, 265: 615-621. An overview of the structure and function of RNA-binding domains.

    PubMed  Google Scholar 

  17. Deo RC, Bonanno JB, Sonenberg N, Burley SK: Recognition of polyadenylate RNA by the poly(A)-binding protein. Cell. 1999, 98: 835-845. A report of the structure of the two amino-terminal RRMs of human PABP bound to poly(A).

    PubMed  Google Scholar 

  18. Adam SA, Nakagawa T, Swanson MS, Woodruff TK, Dreyfuss G: mRNA polyadenylate-binding protein: gene isolation and sequencing and identification of a ribonucleoprotein consensus sequence. Mol Cell Biol. 1986, 6: 2932-2943. This article describes the structure and function of the yeast PAB1 gene.

    PubMed  PubMed Central  Google Scholar 

  19. Fukami-Kobayashi K, Tomoda S, Go M: Evolutionary clustering and functional similarity of RNA-binding proteins. FEBS Lett. 1993, 335: 289-293. 10.1016/0014-5793(93)80749-K. A phylogenetic analysis of RNA-binding domains.

    PubMed  Google Scholar 

  20. Kozlov G, Siddiqui N, Coillet-Matillon S, Trempe JF, Ekiel I, Sprules T, Gehring K: Solution structure of the orphan PABC domain from Saccharomyces cerevisiae poly(A)-binding protein. J Biol Chem. 2002, 277: 22822-22828. 10.1074/jbc.M201230200. A description of the structure of the carboxy-terminal helices of yeast Pab1p and comparison to other PABPC domains.

    PubMed  Google Scholar 

  21. Kozlov G, Trempe JF, Khaleghpour K, Kahvejian A, Ekiel I, Gehring K: Structure and function of the C-terminal PABC domain of human poly(A)-binding protein. Proc Natl Acad Sci USA. 2001, 98: 4409-4413. 10.1073/pnas.071024998. A description of the structure of the carboxy-terminal domain of human PABP and demonstration of its role as a peptide-binding site.

    PubMed  PubMed Central  Google Scholar 

  22. Deo RC, Sonenberg N, Burley SK: X-ray structure of the human hyperplastic discs protein: an ortholog of the C-terminal domain of poly(A)-binding protein. Proc Natl Acad Sci USA. 2001, 98: 4414-4419. 10.1073/pnas.071552198. This article reports the structure of the PABP-related domain within the human HYD protein.

    PubMed  PubMed Central  Google Scholar 

  23. Wahle E, Ruegsegger U: 3'-end processing of pre-mRNA in eukaryotes. FEMS Microbiol Rev. 1999, 23: 277-295. 10.1016/S0168-6445(99)00008-X. A review of mRNA polyadenylation mechanisms.

    PubMed  Google Scholar 

  24. Brais B, Bouchard JP, Xie YG, Rochefort DL, Chretien N, Tome FM, Lafreniere RG, Rommens JM, Uyama E, Nohira O, et al: Short GCG expansions in the PABP2 gene cause oculopharyngeal muscular dystrophy. Nat Genet. 1998, 18: 164-167. A report describing the relationship between OPMD and GCG repeats in the gene encoding PABPN1.

    PubMed  Google Scholar 

  25. Fan X, Rouleau GA: Progress in understanding the pathogenesis of oculopharyngeal muscular dystrophy. Can J Neurol Sci. 2003, 30: 8-14. A review addressing diverse aspects of OPMD pathogenesis, including the relationship of its onset to mutations in the PABPN1 gene.

    PubMed  Google Scholar 

  26. Anderson JT, Wilson SM, Datar KV, Swanson MS: NAB2 : a yeast nuclear polyadenylated RNA-binding protein essential for cell viability. Mol Cell Biol. 1993, 13: 2730-2741. Characterization of the yeast nuclear PABP, Nab2p.

    PubMed  PubMed Central  Google Scholar 

  27. Jacobson A: Poly(A) metabolism and translation: the closed loop model. In Translational Control. Edited by: Hershey JW, Mathews MB, Sonenberg N. 1996, Cold Spring Harbor: Cold Spring Harbor Press, 451-480. A review of poly(A) and PABP function in mRNA translation and decay.

    Google Scholar 

  28. Zhao J, Hyman L, Moore C: Formation of mRNA 3' ends in eukaryotes: mechanism, regulation, and interrelationships with other steps in mRNA synthesis. Microbiol Mol Biol Rev. 1999, 63: 405-445. This article and [29] review mRNA 3'-processing in eukaryotes.

    PubMed  PubMed Central  Google Scholar 

  29. Edmonds M: A history of poly A sequences: from formation to factors to function. Prog Nucleic Acid Res Mol Biol. 2002, 71: 285-389. See [28].

    PubMed  Google Scholar 

  30. Hall TM: Poly(A) tail synthesis and regulation: recent structural insights. Curr Opin Struct Biol. 2002, 12: 82-88. 10.1016/S0959-440X(02)00293-2. A review of specific RNA-binding protein structures essential for mRNA processing and function.

    PubMed  Google Scholar 

  31. Meyer S, Urbanke C, Wahle E: Equilibrium studies on the association of the nuclear poly(A) binding protein with poly(A) of different lengths. Biochemistry. 2002, 41: 6082-6089. 10.1021/bi0160866. Characterization of PABPN1 binding to poly(A).

    PubMed  Google Scholar 

  32. Bienroth S, Keller W, Wahle E: Assembly of a processive messenger RNA polyadenylation complex. EMBO J. 1993, 12: 585-594. The authors of this article and [33] demonstrate that PABPN1 stimulates poly(A) polymerase to switch from distributive to processive synthesis.

    PubMed  PubMed Central  Google Scholar 

  33. Wahle E: Poly(A) tail length control is caused by termination of processive synthesis. J Biol Chem. 1995, 270: 2800-2808. See [32].

    PubMed  Google Scholar 

  34. Keller RW, Kuhn U, Aragon M, Bornikova L, Wahle E, Bear DG: The nuclear poly(A) binding protein, PABP2, forms an oligomeric particle covering the length of the poly(A) tail. J Mol Biol. 2000, 297: 569-583. 10.1006/jmbi.2000.3572. This article reports the binding of PABPN1 monomers to nascent adenylate residues, forming linear filaments and spherical particles.

    PubMed  Google Scholar 

  35. Hector RE, Nykamp KR, Dheur S, Anderson JT, Non PJ, Urbinati CR, Wilson SM, Minvielle-Sebastia L, Swanson MS: Dual requirement for yeast hnRNP Nab2p in mRNA poly(A) tail length control and nuclear export. EMBO J. 2002, 21: 1800-1810. 10.1093/emboj/21.7.1800. Nab2p is shown to function in a manner analogous to PABPN1.

    PubMed  PubMed Central  Google Scholar 

  36. Sachs AB, Davis RW: The poly(A) binding protein is required for poly(A) shortening and 60S ribosomal subunit-dependent translation initiation. Cell. 1989, 58: 857-867. This article reports the phenotypes of yeast suppressor strains harboring deletions of the PAB1 gene.

    PubMed  Google Scholar 

  37. Amrani N, Minet M, Le Gouar M, Lacroute F, Wyers F: Yeast Pab1 interacts with Rna15 and participates in the control of the poly(A) tail length in vitro. Mol Cell Biol. 1997, 17: 3694-3701. Demonstration of the role of Pab1p in yeast mRNA 3'-processing in vitro.

    PubMed  PubMed Central  Google Scholar 

  38. Minvielle-Sebastia L, Preker PJ, Wiederkehr T, Strahm Y, Keller W: The major yeast poly(A)-binding protein is associated with cleavage factor IA and functions in premessenger RNA 3'-end formation. Proc Natl Acad Sci USA. 1997, 94: 7897-7902. 10.1073/pnas.94.15.7897. Demonstration of the role of Pab1p in yeast mRNA 3'-processing.

    PubMed  PubMed Central  Google Scholar 

  39. Zhelkovsky A, Helmling S, Moore C: Processivity of the Saccharomyces cerevisiae poly(A) polymerase requires interactions at the carboxyl-terminal RNA binding domain. Mol Cell Biol. 1998, 18: 5942-5951. The authors demonstrate a regulation of poly(A) polymerase function during yeast mRNA 3'-processing.

    PubMed  PubMed Central  Google Scholar 

  40. Helmling S, Zhelkovsky A, Moore CL: Fip1 regulates the activity of poly(A) polymerase through multiple interactions. Mol Cell Biol. 2001, 21: 2026-2037. 10.1128/MCB.21.6.2026-2037.2001. This article reports the regulation of yeast Pap1p by Fip1p.

    PubMed  PubMed Central  Google Scholar 

  41. Tacahashi Y, Helmling S, Moore CL: Functional dissection of the zinc finger and flanking domains of the Yth1 cleavage/polyadenylation factor. Nucleic Acids Res. 2003, 31: 1744-1752. 10.1093/nar/gkg265. The regulation of yeast mRNA 3'-end formation by Yth1p is demonstrated.

    PubMed  PubMed Central  Google Scholar 

  42. Boeck R, Tarun S, Rieger M, Deardorff JA, Muller-Auer S, Sachs AB: The yeast Pan2 protein is required for poly(A)-binding protein-stimulated poly(A)-nuclease activity. J Biol Chem. 1996, 271: 432-438. 10.1074/jbc.271.1.432. A paper reporting the cloning of yeast PAN2 and its requirement for poly(A) shortening.

    PubMed  Google Scholar 

  43. Brown CE, Sachs AB: Poly(A) tail length control in Saccharomyces cerevisiae occurs by message-specific deadenylation. Mol Cell Biol. 1998, 18: 6548-6559. The authors show that yeast PAN subunits are required for proper poly(A)-tail formation.

    PubMed  PubMed Central  Google Scholar 

  44. Brown CE, Tarun SZ, Boeck R, Sachs AB: PAN3 encodes a subunit of the Pab1p-dependent poly(A) nuclease in Saccharomyces cerevisiae. Mol Cell Biol. 1996, 16: 5744-5753. A paper reporting the cloning of yeast PAN3 and its requirement for mRNA poly(A)-tail shortening.

    PubMed  PubMed Central  Google Scholar 

  45. Mangus DA, Amrani N, Jacobson A: Pbp1p, a factor interacting with Saccharomyces cerevisiae poly(A)-binding protein, regulates polyadenylation. Mol Cell Biol. 1998, 18: 7383-7396. The authors report the cloning of yeast PBP1 and demonstrate its role in mRNA 3'-end formation.

    PubMed  PubMed Central  Google Scholar 

  46. Long RM, Elliott DJ, Stutz F, Rosbash M, Singer RH: Spatial consequences of defective processing of specific yeast mRNAs revealed by fluorescent in situ hybridization. RNA. 1995, 1: 1071-1078. In situ hybridization was used to demonstrate the requirement of polyadenylation for mRNA export.

    PubMed  PubMed Central  Google Scholar 

  47. Huang Y, Carmichael GC: Role of polyadenylation in nucleocy-toplasmic transport of mRNA. Mol Cell Biol. 1996, 16: 1534-1542. Ribozyme constructs were used to demonstrate a relationship between mRNA polyadenylation and export in mammalian cells.

    PubMed  PubMed Central  Google Scholar 

  48. Dower K, Rosbash M: T7 RNA polymerase-directed transcripts are processed in yeast and link 3' end formation to mRNA nuclear export. RNA. 2002, 8: 686-697. 10.1017/S1355838202024068. T7 derived transcripts were used to link polyadenylation and mRNA export in yeast.

    PubMed  PubMed Central  Google Scholar 

  49. Hammell CM, Gross S, Zenklusen D, Heath CV, Stutz F, Moore C, Cole CN: Coupling of termination, 3' processing, and mRNA export. Mol Cell Biol. 2002, 22: 6441-6457. 10.1128/MCB.22.18.6441-6457.2002. This article describes the interrelationships of polyadenylation and export in mutant yeast strains.

    PubMed  PubMed Central  Google Scholar 

  50. Allen NP, Huang L, Burlingame A, Rexach M: Proteomic analysis of nucleoporin interacting proteins. J Biol Chem. 2001, 276: 29268-29274. 10.1074/jbc.M102629200. An article describing the interaction of yeast Pab1p with nucleoporins.

    PubMed  Google Scholar 

  51. Uetz P, Giot L, Cagney G, Mansfield TA, Judson RS, Knight JR, Lockshon D, Narayan V, Srinivasan M, Pochart P, et al: A comprehensive analysis of protein-protein interactions in Saccharomyces cerevisiae. Nature. 2000, 403: 623-627. 10.1038/35001009. This analysis includes the interaction of yeast Pab1p with Xpo1p.

    PubMed  Google Scholar 

  52. Strasser K, Hurt E: Yra1p, a conserved nuclear RNA-binding protein, interacts directly with Mex67p and is required for mRNA export. EMBO J. 2000, 19: 410-420. 10.1093/emboj/19.3.410. The authors describe the function of yeast Yra1p in mRNA export.

    PubMed  PubMed Central  Google Scholar 

  53. Afonina E, Stauber R, Pavlakis GN: The human poly(A)-binding protein 1 shuttles between the nucleus and the cytoplasm. J Biol Chem. 1998, 273: 13015-13021. 10.1074/jbc.273.21.13015. A paper about the nucleocytoplasmic shuttling of human PABPN1.

    PubMed  Google Scholar 

  54. Calado A, Tome FMS, Brais B, Rouleau GA, Kuhn U, Wahle E, Carmo-Fonseca M: Nuclear inclusions in oculopharyngeal muscular dystrophy consist of poly(A)-binding protein 2 aggregates which sequester poly(A) RNA. Hum Mol Genet. 2000, 9: 2321-2328. The authors report the nuclear insolubility of PABPN1 derived from alleles containing GCG repeats.

    PubMed  Google Scholar 

  55. Duncan K, Umen JG, Guthrie C: A putative ubiquitin ligase required for efficient mRNA export differentially affects hnRNP transport. Curr Biol. 2000, 10: 687-696. 10.1016/S0960-9822(00)00527-3. A report including data on the nucleocytoplasmic shuttling of yeast Nab2p.

    PubMed  Google Scholar 

  56. Thakurta AG, Ho Yoon J, Dhar R: Schizosaccharomyces pombe spPABP, a homologue of Saccharomyces cerevisiae Pab1p, is a non-essential, shuttling protein that facilitates mRNA export.Yeast. 2002, 19: 803-810. 10.1002/yea.876. S. pombe PABP shuttles between the nucleus and cytoplasm.,

    PubMed  Google Scholar 

  57. Chen Z, Li Y, Krug RM: Influenza A virus NS1 protein targets poly(A)-binding protein II of the cellular 3'-end processing machinery. EMBO J. 1999, 18: 2273-2283. 10.1093/emboj/18.8.2273. The authors show inhibition of mRNA export by interaction of influenza virus NS1A with PABPN1.

    PubMed  PubMed Central  Google Scholar 

  58. Duvel K, Valerius O, Mangus DA, Jacobson A, Braus GH: Replacement of the yeast TRP4 3' untranslated region by a hammerhead ribozyme results in a stable and efficiently exported mRNA that lacks a poly(A) tail. RNA. 2002, 8: 336-344. 10.1017/S1355838202021039. This article reports that ribozyme-generated mRNAs lacking a poly(A) tail are capable of mRNA export in yeast.

    PubMed  PubMed Central  Google Scholar 

  59. Proweller A, Butler S: Efficient translation of poly(A)- deficient mRNAs in Saccharomyces cerevisiae. Genes Dev. 1994, 8: 2629-2640. mRNAs lacking a poly(A) tail generated in a yeast pap1 mutant enter the cytoplasm.

    PubMed  Google Scholar 

  60. Brodsky AS, Silver PA: Pre-mRNA processing factors are required for nuclear export. RNA. 2000, 6: 1737-1749. 10.1017/S1355838200001059. Mutations in yeast genes regulating mRNA 3'-processing block mRNA export.

    PubMed  PubMed Central  Google Scholar 

  61. Hilleren P, McCarthy T, Rosbash M, Parker R, Jensen TH: Quality control of mRNA 3'-end processing is linked to the nuclear exosome. Nature. 2001, 413: 538-542. 10.1038/35097110. In this article, linkage of mRNA polyadenylation and nuclear export was analyzed in yeast export mutants.

    PubMed  Google Scholar 

  62. Jensen TH, Patricio K, McCarthy T, Rosbash M: A block to mRNA nuclear export in S. cerevisiae leads to hyperadenylation of transcripts that accumulate at the site of transcription. Mol Cell. 2001, 7: 887-898. 10.1016/S1097-2765(01)00232-5. The link between mRNA 3'-processing and nuclear export was analyzed in yeast export mutants.

    PubMed  Google Scholar 

  63. Jacobson A, Favreau M: Possible involvement of poly(A) in protein synthesis. Nucleic Acids Res. 1983, 11: 6353-6368. This is the first formulation of the closed-loop model postulating interactions between the 5' and 3' ends of the mRNA.

    PubMed  PubMed Central  Google Scholar 

  64. Tarun SZ, Sachs AB: Association of the yeast poly(A) tail binding protein with translation initiation factor eIF-4G. EMBO J. 1996, 15: 7168-7177. Demonstration that yeast Pab1p interacts with eIF4G.

    PubMed  PubMed Central  Google Scholar 

  65. Tarun SZ, Wells SE, Deardorff JA, Sachs AB: Translation initiation factor eIF4G mediates in vitro poly(A) tail-dependent translation. Proc Natl Acad Sci USA. 1997, 94: 9046-9051. 10.1073/pnas.94.17.9046. Further characterization of eIF4G/Pab1p interaction in yeast.

    PubMed  PubMed Central  Google Scholar 

  66. Wells SE, Hillner PE, Vale RD, Sachs AB: Circularization of mRNA by eukaryotic translation initiation factors. Mol Cell. 1998, 2: 135-140. This article reports the in vitro formation of mRNA closed loops in the presence of Pab1p, eIF4G, and eIF4E.

    PubMed  Google Scholar 

  67. Munroe D, Jacobson A: mRNA poly(A) tail, a 3' enhancer of translational initiation. Mol Cell Biol. 1990, 10: 3441-3455. This article shows the stimulation of 60S ribosome joining by the presence of an mRNA poly(A) tail.

    PubMed  PubMed Central  Google Scholar 

  68. Gallie DR: The cap and poly(A) tail function synergistically to regulate mRNA translational efficiency. Genes Dev. 1991, 5: 2108-2116. Synergistic translational effects of mRNA cap and poly(A) tail are reported.

    PubMed  Google Scholar 

  69. Haghighat A, Sonenberg N: eIF4G dramatically enhances the binding of eIF4E to the mRNA 5'-cap structure. J Biol Chem. 1997, 272: 21677-21680. 10.1074/jbc.272.35.21677. The authors show the effects of eIF4G-binding on the interaction of the initiation factor eIF4E with the mRNA cap.

    PubMed  Google Scholar 

  70. Ptushkina M, von der Haar T, Vasilescu S, Frank R, Birkenhager R, McCarthy JE: Cooperative modulation by eIF4G of eIF4E-binding to the mRNA 5' cap in yeast involves a site partially shared by p20. EMBO J. 1998, 17: 4798-4808. 10.1093/emboj/17.16.4798. This work shows that eIF4G/eIF4E interaction enhances eIF4E affinity for the cap.

    PubMed  PubMed Central  Google Scholar 

  71. Borman AM, Michel YM, Kean KM: Biochemical characterisation of cap-poly(A) synergy in rabbit reticulocyte lysates: the eIF4G-PABP interaction increases the functional affinity of eIF4E for the capped mRNA 5'-end. Nucleic Acids Res. 2000, 28: 4068-4075. 10.1093/nar/28.21.4068. This article shows that PABP/eIF4G interaction enhances eIF4E-binding of the mRNA cap.

    PubMed  PubMed Central  Google Scholar 

  72. Luo Y, Goss DJ: Homeostasis in mRNA initiation: wheat germ poly(A)-binding protein lowers the activation energy barrier to initiation complex formation. J Biol Chem. 2001, 276: 43083-43086. 10.1074/jbc.M104970200. The authors report that the dissociation of the wheat germ cap-binding complex from the mRNA cap is slowed by interaction with PABP.

    PubMed  Google Scholar 

  73. Le H, Tanguay RL, Balasta ML, Wei CC, Browning KS, Metz AM, Goss DJ, Gallie DR: Translation initiation factors eIF-iso4G and eIF-4B interact with the poly(A)-binding protein and increase its RNA binding activity. J Biol Chem. 1997, 272: 16247-16255. 10.1074/jbc.272.26.16247. PABP affinity for mRNA is shown to be enhanced by interaction with the cap-binding complex and eIF4B.

    PubMed  Google Scholar 

  74. Bi X, Goss DJ: Wheat germ poly(A)-binding protein increases the ATPase and the RNA helicase activity of translation initiation factors eIF4A, eIF4B, and eIF-iso4F. J Biol Chem. 2000, 275: 17740-17746. 10.1074/jbc.M909464199. Interaction with plant PABP is shown to enhance initiation factor activities.

    PubMed  Google Scholar 

  75. Sachs A: Physical and functional interactions between the mRNA cap structure and the poly(A) tail. In Translational Control of Gene Expression. Edited by: Sonenberg N, Hershey JW, Mathews MB. 2000, Cold Spring Harbor: Cold Spring Harbor Laboratory Press, 447-465. A review of interactions between the 5' and 3' ends of mRNPs.

    Google Scholar 

  76. Kessler SH, Sachs AB: RNA recognition motif 2 of yeast Pab1p is required for its functional interaction with eukaryotic translation initiation factor 4G. Mol Cell Biol. 1998, 18: 51-57. The authors report the identification of RRM2 as the PABP domain interacting with eIF4G.

    PubMed  PubMed Central  Google Scholar 

  77. Imataka H, Gradi A, Sonenberg N: A newly identified N-terminal amino acid sequence of human eIF4G binds poly(A)-binding protein and functions in poly(A)-dependent translation. EMBO J. 1998, 17: 7480-7489. 10.1093/emboj/17.24.7480. This article describes the eIF4G/PABP interaction in mammalian cells.

    PubMed  PubMed Central  Google Scholar 

  78. Otero LJ, Ashe MP, Sachs AB: The yeast poly(A)-binding protein Pab1p stimulates in vitro poly(A)-dependent and cap-dependent translation by distinct mechanisms. EMBO J. 1999, 18: 3153-3163. 10.1093/emboj/18.11.3153. A paper reporting the identification of distinct Pab1p domains required for translation-stimulation activities.

    PubMed  PubMed Central  Google Scholar 

  79. Craig AW, Haghighat A, Yu AT, Sonenberg N: Interaction of polyadenylate-binding protein with the eIF4G homologue PAIP enhances translation. Nature. 1998, 392: 520-523. 10.1038/33198. The identification and characterization of PAIP-1, a mammalian PABP-interacting protein.

    PubMed  Google Scholar 

  80. Roy G, De Crescenzo G, Khaleghpour K, Kahvejian A, O'Connor-McCourt M, Sonenberg N: Paip1 interacts with poly(A)-binding protein through two independent binding motifs. Mol Cell Biol. 2002, 22: 3769-3782. 10.1128/MCB.22.11.3769-3782.2002. The identification of PAIP1 sites that interact with PABP.

    PubMed  PubMed Central  Google Scholar 

  81. Khaleghpour K, Svitkin YV, Craig AW, DeMaria CT, Deo RC, Burley SK, Sonenberg N: Translational repression by a novel partner of human poly(A)-binding protein, Paip2. Mol Cell. 2001, 7: 205-216. The identification and characterization of Paip2, a PABP interacting protein that inhibits translation.

    PubMed  Google Scholar 

  82. Khaleghpour K, Kahvejian A, De Crescenzo G, Roy G, Svitkin YV, Imataka H, O'Connor-McCourt M, Sonenberg N: Dual interactions of the translational repressor Paip2 with poly(A)-binding protein. Mol Cell Biol. 2001, 21: 5200-5213. 10.1128/MCB.21.15.5200-5213.2001. The identification and characterization of interaction domains in PABP and Paip2.

    PubMed  PubMed Central  Google Scholar 

  83. Gray NK, Coller JM, Dickson KS, Wickens M: Multiple portions of poly(A)-binding protein stimulate translation in vivo. EMBO J. 2000, 19: 4723-4733. 10.1093/emboj/19.17.4723. Tethered PABP was used to define PABP domains required for translational stimulation and mRNA stabilization in Xenopus oocytes and yeast.

    PubMed  PubMed Central  Google Scholar 

  84. Sachs AB, Varani G: Eukaryotic translation initiation: there are (at least) two sides to every story. Nat Struct Biol. 2000, 7: 356-361. 10.1038/75120. A review of PABP structures required for interaction with eIF4G and poly(A).

    PubMed  Google Scholar 

  85. Searfoss A, Dever TE, Wickner R: Linking the 3' poly(A) tail to the subunit joining step of translation initiation: relations of Pab1p, eukaryotic translation initiation factor 5b (Fun12p), and Ski2p-Slh1p. Mol Cell Biol. 2001, 21: 4900-4908. 10.1128/MCB.21.15.4900-4908.2001. Analyses of yeast mutants implicating PABP in the regulation of 60S subunit joining.

    PubMed  PubMed Central  Google Scholar 

  86. Groft CM, Burley SK: Recognition of eIF4G by rotavirus NSP3 reveals a basis for mRNA circularization. Mol Cell. 2002, 9: 1273-1283. The authors report that rotavirus NSP3 functions as a PABP analog.

    PubMed  Google Scholar 

  87. Lamphear BJ, Kirchweger R, Skern T, Rhoads RE: Mapping of functional domains in eukaryotic protein synthesis initiation factor 4G (eIF4G) with picornaviral proteases. Implications for cap-dependent and cap-independent translational initiation. J Biol Chem. 1995, 270: 21975-21983. 10.1074/jbc.270.37.21975. An article describing the cleavage of PABP and eIF4G by viral proteases.

    PubMed  Google Scholar 

  88. Kuyumcu-Martinez NM, Joachims M, Lloyd RE: Efficient cleavage of ribosome-associated poly(A)-binding protein by enterovirus 3C protease. J Virol. 2002, 76: 2062-2074. 10.1128/jvi.76.5.2062-2074.2002. An article describing the proteolytic cleavage of PABP during viral infection.

    PubMed  PubMed Central  Google Scholar 

  89. Le H, Browning KS, Gallie DR: The phosphorylation state of poly(A)-binding protein specifies its binding to poly(A) RNA and its interaction with eukaryotic initiation factor (eIF) 4F, eIFiso4F, and eIF4B. J Biol Chem. 2000, 275: 17452-17462. 10.1074/jbc.M001186200. The authors report that the activity of plant PABP is regulated by phosphorylation.

    PubMed  Google Scholar 

  90. Lee J, Bedford MT: PABP1 identified as an arginine methyltransferase substrate using high-density protein arrays. EMBO Rep. 2002, 3: 268-273. 10.1093/embo-reports/kvf052. An article describing the mapping of a site of arginine methylation in PABP.

    PubMed  PubMed Central  Google Scholar 

  91. Hornstein E, Git A, Braunstein I, Avni D, Meyuhas O: The expression of poly(A)-binding protein gene is translationally regulated in a growth-dependent fashion through a 5'-terminal oligopyrimidine tract motif. J Biol Chem. 1999, 274: 1708-1714. 10.1074/jbc.274.3.1708. The authors report that regulation of PABP expression is mediated by a 5'-terminal oligopyrimidine tract motif.

    PubMed  Google Scholar 

  92. Bag J: Feedback inhibition of poly(A)-binding protein mRNA translation. A possible mechanism of translation arrest by stalled 40S ribosomal subunits. J Biol Chem. 2001, 276: 47352-47360. 10.1074/jbc.M107676200. The authors shows autogenous regulation of PABP mRNA translation in reticulocyte lysates.

    PubMed  Google Scholar 

  93. Mazumder B, Seshadri V, Imataka H, Sonenberg N, Fox PL: Translational silencing of ceruloplasmin requires the essential elements of mRNA circularization: poly(A) tail, poly(A)-binding protein, and eukaryotic translation initiation factor 4G. Mol Cell Biol. 2001, 21: 6440-6449. 10.1128/MCB.21.19.6440-6449.2001. PABP is shown to be a facilitator of translational repression.

    PubMed  PubMed Central  Google Scholar 

  94. Richter JD: Dynamics of poly(A) addition and removal during development. In Translational Control. Edited by: Hershey JW, Mathews MB, Sonenberg N. 1996, Cold Spring Harbor: Cold Spring Harbor Press, 481-503. A review of the role of PABP in developmentally regulated mRNA translation.

    Google Scholar 

  95. Yohn CB, Cohen A, Danon A, Mayfield SP: A poly(A)-binding protein functions in the chloroplast as a message-specific translation factor. Proc Natl Acad Sci USA. 1998, 95: 2238-2243. 10.1073/pnas.95.5.2238. PABP is shown to function as an mRNA-specific translational activator in C. reinhardtii.

    PubMed  PubMed Central  Google Scholar 

  96. Kisselev L, Ehrenberg M, Frolova L: Termination of translation: interplay of mRNA, rRNAs and release factors?. EMBO J. 2003, 22: 175-182. 10.1093/emboj/cdg017. A review of translation termination mechanisms and factors.

    PubMed  PubMed Central  Google Scholar 

  97. 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. Direct association of erf3/GSPT with polyadenylate-binding protein. J Biol Chem. 1999, 274: 16677-16680. 10.1074/jbc.274.24.16677. This article reports the interactions of mammalian PABP and the termination factor eRF3.

    PubMed  Google Scholar 

  98. 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. 10.1128/MCB.22.10.3301-3315.2002. This article reports the interactions of yeast PABP and termination factor eRF3.

    PubMed  PubMed Central  Google Scholar 

  99. Uchida N, Hoshino S, Imataka H, Sonenberg N, Katada T: A novel role of the mammalian GSPT/eRF3 associating with poly(A)-binding protein in Cap/Poly(A)-dependent translation. J Biol Chem. 2002, 277: 50286-50292. 10.1074/jbc.M203029200. A possible role of interaction of PABP and eRF3 in ribosome recycling is reported.

    PubMed  Google Scholar 

  100. Jacobson A, Peltz SW: Destabilization of nonsense-containing transcripts in Saccharomyces cerevisiae. In Translational Control of Gene Expression. Edited by: Sonenberg N, Hershey JWB, Mathews MB. 2000, Cold Spring Harbor: Cold Spring Harbor Laboratory Press, 827-847. A review of factors, sequences, and mechanisms regulating nonsense-mediated mRNA decay in yeast.

    Google Scholar 

  101. Jacobson A, Peltz SW: Interrelationships of the pathways of mRNA decay and translation in eukaryotic cells. Annu Rev Biochem. 1996, 65: 693-739. 10.1146/annurev.bi.65.070196.003401. A review of the role of mRNP 5'/3' interactions in the regulation of eukaryotic mRNA stability and translation.

    PubMed  Google Scholar 

  102. Decker CJ, Parker R: mRNA decay enzymes: decappers conserved between yeast and mammals. Proc Natl Acad Sci USA. 2002, 99: 12512-12514. 10.1073/pnas.212518099. A review of mRNA-decay mechanisms in yeast and mammalian cells.

    PubMed  PubMed Central  Google Scholar 

  103. Schwartz DC, Parker R: Mutations in translation initiation factors lead to increased rates of deadenylation and decapping of mRNAs in Saccharomyces cerevisiae. Mol Cell Biol. 1999, 19: 5247-5256. The article reports a relationship between mRNA translation initiation and decay.

    PubMed  PubMed Central  Google Scholar 

  104. Schwartz DC, Parker R: mRNA decapping in yeast requires dissociation of the cap binding protein, eukaryotic translation initiation factor 4E. Mol Cell Biol. 2000, 20: 7933-7942. 10.1128/MCB.20.21.7933-7942.2000. Yeast experiments indicate that eIF4E dissociation from the cap is required for mRNA decapping.

    PubMed  PubMed Central  Google Scholar 

  105. Chen CY, Xu N, Shyu AB: mRNA decay mediated by two distinct AU-rich elements from c-fos and granulocyte-macrophage colony-stimulating factor transcripts: different deadenylation kinetics and uncoupling from translation. Mol Cell Biol. 1995, 15: 5777-5788. A report of the relationships between poly(A) shortening and decay in mammalian cells.

    PubMed  PubMed Central  Google Scholar 

  106. Muhlrad D, Decker CJ, Parker R: Turnover mechanisms of the stable yeast PGK1 mRNA. Mol Cell Biol. 1995, 15: 2145-2156. This work shows that poly(A) shortening is not the rate-limiting step in the decay of the yeast PGK1 mRNA.

    PubMed  PubMed Central  Google Scholar 

  107. Caponigro G, Parker R: Multiple functions for the poly(A)-binding protein in mRNA decapping and deadenylation in yeast. Genes Dev. 1995, 9: 2421-2432. The authors of this article and [108] describe the effect of yeast pab1 mutations on mRNA decay.

    PubMed  Google Scholar 

  108. Morrissey JP, Deardorff JA, Hebron C, Sachs AB: Decapping of stabilized, polyadenylated mRNA in yeast pab1 mutants. Yeast. 1999, 15: 687-702. 10.1002/(SICI)1097-0061(19990615)15:8<687::AID-YEA412>3.3.CO;2-C. See [107].

    PubMed  Google Scholar 

  109. Coller JM, Gray NK, Wickens MP: mRNA stabilization by poly(A) binding protein is independent of poly(A) and requires translation. Genes Dev. 1998, 12: 3226-3235. An article showing stabilization of poly(A)- yeast mRNA by tethered Pab1p.

    PubMed  PubMed Central  Google Scholar 

  110. Bernstein P, Peltz SW, Ross J: The poly(A)-poly(A)-binding protein complex is a major determinant of mRNA stability in vitro. Mol Cell Biol. 1989, 9: 659-670. The authors show that cell-free extracts depleted of PABP enhance mRNA-decay rates.

    PubMed  PubMed Central  Google Scholar 

  111. Ford LP, Bagga PS, Wilusz J: The poly(A) tail inhibits the assembly of a 3'-to-5' exonuclease in an in vitro RNA stability system. Mol Cell Biol. 1997, 17: 398-406. This article reports PABP inhibition of mRNA 3' to 5' decay in vitro.

    PubMed  PubMed Central  Google Scholar 

  112. Wilusz CJ, Gao M, Jones CL, Wilusz J, Peltz SW: Poly(A)-binding proteins regulate both mRNA deadenylation and decapping in yeast cytoplasmic extracts. RNA. 2001, 7: 1416-1424. Cell-free decay of yeast mRNAs is shown to be inhibited by PABPs.

    PubMed  PubMed Central  Google Scholar 

  113. Wang Z, Day N, Trifillis P, Kiledjian M: An mRNA stability complex functions with poly(A)-binding protein to stabilize mRNA in vitro. Mol Cell Biol. 1999, 19: 4552-4560. This article reports that cell-free deadenylation of mammalian mRNAs occurs incrementally.

    PubMed  PubMed Central  Google Scholar 

  114. Korner CG, Wormington M, Muckenthaler M, Schneider S, Dehlin E, Wahle E: The deadenylating nuclease (DAN) is involved in poly(A) tail removal during the meiotic maturation of Xenopus oocytes. EMBO J. 1998, 17: 5427-5437. 10.1093/emboj/17.18.5427. Mammalian deadenylase is shown to be inhibited by PABP.

    PubMed  PubMed Central  Google Scholar 

  115. Wang Z, Kiledjian M: The poly(A)-binding protein and an mRNA stability protein jointly regulate an endoribonuclease activity. Mol Cell Biol. 2000, 20: 6334-6341. 10.1128/MCB.20.17.6334-6341.2000. PABP stimulation of α3CP protein binding antagonizes endonuclease ErEN.

    PubMed  PubMed Central  Google Scholar 

  116. Shyu AB, Belasco JG, Greenberg ME: Two distinct destabilizing elements in the c-fos message trigger deadenylation as a first step in rapid mRNA decay. Genes Dev. 1991, 5: 221-231. Coding region and 3' UTR elements are shown to destabilize the c-fos mRNA.

    PubMed  Google Scholar 

  117. Grosset C, Chen CY, Xu N, Sonenberg N, Jacquemin-Sablon H, Shyu AB: A mechanism for translationally coupled mRNA turnover: interaction between the poly(A) tail and a c-fos RNA coding determinant via a protein complex. Cell. 2000, 103: 29-40. A stability element in the coding region of the c-fos mRNA interacts with a complex of proteins that includes PABP.

    PubMed  Google Scholar 

  118. Tharun S, Parker R: Targeting an mRNA for decapping: displacement of translation factors and association of the Lsm1p-7p complex on deadenylated yeast mRNAs. Mol Cell. 2001, 8: 1075-1083. Deadenylation of yeast mRNAs results in the binding of Lsm1p-7p complex.

    PubMed  Google Scholar 

  119. Sheth U, Parker R: Decapping and decay of messenger RNA occur in cytoplasmic processing bodies. Science. 2003, 300: 805-808. 10.1126/science.1082320. P bodies are reported to be possible sites of mRNA decay in yeast.

    PubMed  PubMed Central  Google Scholar 

  120. Chekanova JA, Shaw RJ, Belostotsky DA: Analysis of an essential requirement for the poly(A) binding protein function using cross-species complementation. Curr Biol. 2001, 11: 1207-1214. 10.1016/S0960-9822(01)00371-2. This article reports the complementation of yeast pab1 mutation by Arabidopsis PAB3.

    PubMed  Google Scholar 

  121. Wang X, Ullah Z, Grumet R: Interaction between zucchini yellow mosaic potyvirus RNA-dependent RNA polymerase and host poly-(A) binding protein. Virology. 2000, 275: 433-443. 10.1006/viro.2000.0509. PABP is shown to be a subunit of a plant viral replicase.

    PubMed  Google Scholar 

  122. Thompson JD, Higgins DG, Gibson TJ: CLUSTAL W: improving the sensitivity of progressive multiple sequence alignment through sequence weighting, position-specific gap penalties and weight matrix choice. Nucleic Acids Res. 1994, 22: 4673-4680. Description of a program that can be used to generate multiple sequence alignments.

    PubMed  PubMed Central  Google Scholar 

  123. Saitou N, Nei M: The neighbor-joining method: a new method for reconstructing phylogenetic trees. Mol Biol Evol. 1987, 4: 406-425. A methodology for phylogenetic analysis.

    PubMed  Google Scholar 

  124. Phylodendron. A website with a application for drawing phylogenetic trees., [http://iubio.bio.indiana.edu/treeapp/phylodendron-doc.html]

  125. Merritt EAfs, Bacon DJ: Raster 3D: photorealistic molecular graphics. In Methods in Enzymol. Edited by: Carter Jr CW, Sweet RM. 1997, New York: Academic Press, 505-524. This article and [126] describe software for molecular graphics.

    Google Scholar 

  126. Kraulis PJ: MOLSCRIPT: a program to produce both detailed and schematic plots of protein structures. J Appl Crystallogr. 1991, 24: 946-950. 10.1107/S0021889891004399.

    Google Scholar 

Download references

Acknowledgements

This work was supported by grants GM27757 and GM61096 to A.J. from the National Institutes of Health. We thank Dmitry Belostotsky for sharing data prior to publication, Bin Qin for invaluable assistance with computer graphics, and members of our laboratory for comments on the manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Allan Jacobson.

Electronic supplementary material

13059_2003_637_MOESM1_ESM.doc

Additional data file 1: A list of the currently known PABP genes with accession numbers and links to their entries in the nucleotide and protein sequence databases (DOC 140 KB)

Additional data file 2: The sequences of the currently known PABP proteins in FASTA format (FASTA 78 KB)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Mangus, D.A., Evans, M.C. & Jacobson, A. Poly(A)-binding proteins: multifunctional scaffolds for the post-transcriptional control of gene expression. Genome Biol 4, 223 (2003). https://doi.org/10.1186/gb-2003-4-7-223

Download citation

  • Published:

  • DOI: https://doi.org/10.1186/gb-2003-4-7-223

Keywords