Skip to main content

The retrovirus RNA trafficking granule: from birth to maturity

Abstract

Post-transcriptional events in the life of an RNA including RNA processing, transport, translation and metabolism are characterized by the regulated assembly of multiple ribonucleoprotein (RNP) complexes. At each of these steps, there is the engagement and disengagement of RNA-binding proteins until the RNA reaches its final destination. For retroviral genomic RNA, the final destination is the capsid. Numerous studies have provided crucial information about these processes and serve as the basis for studies on the intracellular fate of retroviral RNA. Retroviral RNAs are like cellular mRNAs but their processing is more tightly regulated by multiple cis-acting sequences and the activities of many trans-acting proteins. This review describes the viral and cellular partners that retroviral RNA encounters during its maturation that begins in the nucleus, focusing on important events including splicing, 3' end-processing, RNA trafficking from the nucleus to the cytoplasm and finally, mechanisms that lead to its compartmentalization into progeny virions.

Background

The life of an mRNA is directed by the protein components of ribonucleoprotein particles (RNP) whose roles include nuclear processing reactions, transport, translation and degradation. Retroviral replication depends on many of the same processes to form viral mRNA and genomic RNA providing an experimentally tractable system to study the cis and trans determinants of mRNA fate. In this review, we summarize the current understanding of the processes affecting retroviral RNA metabolism as the RNA moves from its site of synthesis within the nucleus to its encapsidation into viral particles that emerge from the plasma membrane. The field has not only illuminated the cellular processes regulating RNA fate in general but also provided insights into potential strategies to impair replication of these viral pathogens.

Preserving genome-length RNA

Splicing control – the role of the NRS of ASV

The expression of viral proteins from unspliced, incompletely spliced and fully spliced transcripts has necessitated that retroviruses evolve strategies to control the extent of RNA splicing. Extensive studies of avian sarcoma virus (ASV) splicing revealed three mechanisms of splicing control. The first involves the maintenance of suboptimal 3' splice site (ss) signals. Use of the env 3'ss is controlled by a suboptimal branchpoint (bpt) sequence and a nearby exonic splicing enhancer (ESE) [1, 2] whereas the src 3' ss has a suboptimal pyrimidine tract (ppt) [3]. Mutations that improved the quality of the signals increased splicing and had detrimental effects on replication. Consistent with a requirement of inefficient splicing for optimal replication, revertants contained mutations that restored inefficient splicing. In addition to suboptimal splicing signals, a second, poorly characterized negative element is also present upstream of the src 3' ss [4, 5]. Whether this element represents an intronic splicing silencer (ISS) and what factors bind to it remains to be determined. These two splicing control mechanisms are shared with HIV (discussed below). A third, novel control element in Rous sarcoma virus (RSV), that is apparently unique to avian retroviruses, is the negative regulator of splicing, or NRS [6, 7]. The NRS is thought to represent an elaborate pseudo-5'ss that non-productively interacts with and sequesters the viral 3' splice sites such that productive splicing with the authentic 5'ss cannot occur (Figure 1). In addition to its established role in splicing control, the NRS serves a second function in promoting efficient polyadenylation of viral transcripts, as discussed below.

Figure 1
figure 1

Model for NRS effects on splicing and polyadenylation. Schematic of RSV RNA with exons depicted as boxes and introns shown as thin lines. The light shading represents the upstream SR protein binding region of the bipartite NRS, and the darker shading depicts the region that binds U1 snRNP. SR proteins promote U1 binding, which initiates early interactions with factors associated with the viral 3' splice site (env in this example), and this is thought to mature into a spliceosome-like NRS inhibitory complex (indicated by the large oval) that forms between the NRS and the viral 3' splice site but which is catalytically inactive (an X over the intron); a possible role for U6 snRNP is indicated by the question mark. The NRS complex sequesters the 3'ss from interacting with the authentic viral 5'ss to block splicing. The NRS complex may influence polyadenylation by serving to stabilize the binding of splicing factors to the weak viral 3' splice site, which can then either recruit or stabilize the polyadenylation complex (arrow) and thereby enhance polyadenylation of viral unspliced RNA. U11 snRNP modulates NRS function by antagonizing U1 binding and assembly of the NRS inhibitory complex. A downstream region (intermediate shading) binds hnRNP H, which recruits U11 to a site that overlaps the U1 binding site.

The NRS was originally identified from gag intron deletion mutations that increased splicing in RSV [6–8]. The ability of the NRS to block splicing of heterologous introns facilitated elucidation of factors that bind to it and its mechanism of action [6]. Unlike other negative elements in HIV and RSV that are close to 3' splice sites, the ~230 nt NRS is located in the gag intron approximately 300 nt from the 5'ss and more than 4000 nt from the first of two alternative 3' splice sites (env and src) [9]. Mutagenesis studies determined the NRS to be bipartite; splicing inhibition requires a diffuse upstream purine-rich element separated by ~115 nt from a discrete downstream sequence that resembles a conventional but degenerate U1-type 5' ss [9–12]. Overlapping the degenerate U1-type 5'ss is a consensus binding site for U11 snRNP, a factor that serves an analogous role to U1 in binding the 5'ss of a rare class of introns that are spliced by a second, low abundance spliceosome [13]. Binding of both U1 and U11 to the NRS has been demonstrated however it is the interaction with U1 that leads to splicing inhibition [10, 11, 14]. The mechanism by which U1 binding to the NRS leads to inhibition rather than NRS splicing is not clear, but may involve an aberrant U6 interaction at a later step (M.T.M., unpublished). The U1/U11 sites overlap and thus binding is mutually exclusive. U11 binding may regulate splicing inhibition by modulating U1 binding, and contribute to the balance of unspliced to spliced RNA and replication. Thus, determining the cis and trans factors that govern U1 and U11 binding was important.

The upstream, purine-rich region of the NRS was shown to have potent splicing enhancer activity and to bind members of the SR protein family of splicing factors and hnRNP H [11, 15, 16]. One function of splicing enhancers and SR proteins is recruitment of components of the splicing apparatus. In the case of the NRS, it was shown that the role of the enhancer region and SR proteins was to recruit U1 to the downstream degenerate 5'ss. In contrast, the SR protein-binding region was not necessary for efficient U11 binding [11]. The NRS itself forms an early spliceosome-like complex that is dependent on U1 and SR proteins, and this complex can interact with a 3'ss in a U1-dependent manner [17–19]. This interaction persists into an ATP-dependent, more mature splicing-like complex, however this complex is distinguished from authentic splicing complexes in that the U4:U6/U5 tri-snRNP is not stably bound and the U5-associated protein Prp8 cannot be cross-linked to the 5' ss [198]. It is this aberrant snRNP association that presumably accounts for assembly of a non-catalytic complex that leads to sequestration of the viral 3' ss and culminates in splicing inhibition.

The determinants for U11 binding are largely distinct from U1. U11 is at a competitive disadvantage for NRS binding, being 100-fold less abundant than U1 [20]. It was recently shown that optimal U11 binding requires an upstream 3'ss-like sequence and a downstream G-rich region [21]. The downstream G-rich region binds hnRNP H, and mutations in the G-tracks or depletion of hnRNP H reduces U11 binding in vitro and in vivo [22]. These lessons from RSV suggested a more general role for hnRNP H in U11 binding and splicing of authentic minor-class introns. Indeed, the SCN4A and P120 minor-class introns have G tracts, bind hnRNP H, and require hnRNP H for optimal splicing [22]. HnRNP H also plays a role in U1 binding to an HIV-1 enhancer [23], which is consistent with recent demonstrations that splicing of some U2-dependent introns requires hnRNP H [24, 25].

HIV-1 splicing

In contrast to murine leukemia and avian sarcoma viruses, the increased coding capacity of HIV-1 has necessitated the evolution of a more complex splicing regimen. In addition to structural proteins, HIV-1 expresses six additional proteins that regulate various facets of the virus lifecycle [26]. To account for this increased coding potential, the 9 kb HIV-1 transcript is processed into over 30 mRNAs through alternative splicing [27, 28]. The products are grouped into three size classes: the unspliced, 9 kb RNA encoding Gag and Gag/Pol, the 4 kb, singly spliced RNAs that encode Vif, Vpr, Vpu and Env, and the 2 kb, multiply spliced RNAs that express Tat, Rev and Nef. Generation of the required viral RNAs is achieved through the combinatorial use of five 5' splice sites (SD1-5) and nine 3' splice sites (ss) (SA1-3, SA4a,b,c, SA5-SA7). The production of a spectrum of RNAs from unspliced to multiply spliced necessitated the development of multiple mechanisms to control the extent of viral RNA splicing since a substantial amount of unspliced RNA is needed for replication. Initial analysis of HIV-1 RNA processing focused on the splice sites themselves and demonstrated that while the 5'ss were highly active, the 3'ss were suboptimal due to alterations in either the ppt or bpt sequences [29–31] (Figure 2).

Figure 2
figure 2

Processing of HIV-1 RNA. Outlined in the figure are the cis-acting components of the HIV-1 RNA which control its processing. Indicated are the positions of the 5' splice sites (arrows above the unspliced RNA), 3' splice sites (brackets below the unspliced RNA), and the various ESS and ESE elements that modulate splice site use. At top is an outline of viral genome and on the bottom, the exons which comprise the major spliced forms (4.0 kb singly spliced and 1.8 kb multiply spliced) of the genomic RNA are indicated by black boxes. Multiple spliced RNAs combining or excluding various exons encode each of the viral accessory proteins.

Subsequent research determined that the suboptimal nature of the 3'ss of HIV-1 was not the only point of regulation. It was established that exon sequences also influence the use of individual 3'ss. These exon regulatory sequences fall into two groups; exon splicing enhancers (ESEs) that act to enhance recognition and use of the adjacent splice site, and exon splicing silencers (ESSs) that suppress the use of adjacent 3'ss. To date, four ESSs have been mapped and control the use of the 3'ss for Vpr (ESS-V), Tat (ESS2, ESS2p), and the terminal 3'ss (ESS3) [30, 32–37]. For ESS-V, ESS2, and ESS3, function is dependent upon an interaction with members of the hnRNP A/B protein family [34, 38–40] that results in an early block to spliceosome formation. In the case of ESS3, initial work suggested that binding of hnRNP A1 to ESS3 initiates oligomerization of hnRNP A1 along the RNA, sterically hindering recognition of the ppt and bpt sites by the corresponding splicing factors [41]. Subsequent analyses suggested an alternative mechanism and the involvement of an intronic splicing silencer (ISS) to which hnRNP A1 also binds [42, 43]. Multiple hnRNP A1 binding sites have also been mapped within ESS3 [42]. Mutations that disrupt hnRNP A1 binding to either the ISS or ESS3 result in partial alleviation of inhibition and mutation of both is more severe [40], suggesting hnRNP A1 proteins, bound at ESS3 and ISS, might interact to loop out the intervening sequence and impair splicing factor binding to the bpt and ppt. Such a looping mechanism involving hnRNP A1 binding to separate sites has been proposed for regulation of exon 7B in the hnRNP A1 pre-mRNA [44, 45]. Function of ESS2p is less well studied but correlates with hnRNP H binding [35].

Countering the inhibitory signals of the ESSs are the three ESEs present within the first (ESE2, GAR) and second (ESE3) coding exon of Tat [32, 37, 46–50]. Through interaction with one of several members of the SR protein family, the ESEs act by facilitating the recruitment to and/or stabilization of factors that bind the adjacent 3'ss [51–54]. Overexpression of SF2/ASF leads to enhanced use of SA2 and to a lesser extent SA1 [55, 56], and increased expression of SC35 and SRp40 augment use of SA3, presumably by blocking hnRNP A1 binding to the adjacent ESS2 [55, 56]. A similar competition model was suggested to explain the countering activities of ESE3 and ESS3 that affect SA7 use [42, 48, 49]. In contrast to ESE2 and ESE3, the enhancer downstream of SA5 (GAR) appears to be more complicated. The 5' portion of this bipartite ESE binds SF2/ASF and the 3' half interacts with SRp40 [46]. Point mutations within GAR that abrogate factor binding to either domain reduce the efficiency of this element. In addition to promoting SA5 use, this ESE also functions in the recognition of the downstream 5'ss (SD4) by U1 snRNP. Inactivation of the GAR enhancer results in a dramatic increase in the ligation of SD1 to SA7, bypassing all of the splice acceptors used to produce the viral regulatory protein mRNAs (SA3, SA4a-c, SA5) [46]. Therefore, this GAR enhancer appears to play a critical role in ensuring the correct processing of HIV-1 RNA. As one indication of the delicate balance required to achieve the needed levels of the various viral RNAs, a point mutation within env results in generation of aberrant spliced products due to the creation of a splicing enhancer that activates a cryptic 3'ss (SA6) [23, 57].

Control of splicing in other retroviruses

While RNA processing has been most extensively studied in ASV and HIV-1, work in other systems has also illuminated patterns of splicing regulation. Studies of equine infectious anemia virus (EIAV) have identified both cis and trans modulators of RNA processing. Examination of EIAV splice sites revealed that both the 5'ss and bpt sequences do not deviate significantly from consensus. In contrast, the polypyrimidine tracts are interrupted by purines, which may reduce splice site usage by decreasing binding of the splicing factor U2AF [58]. The purine-rich element (PRE) that comprises the EIAV Rev (eRev) binding site is also involved in splicing regulation [59, 60]. Deletion of the PRE results in a marked increase in unspliced and Tat-encoding RNAs and a reduction in eRev RNA [58]. This observation suggests that the PRE acts like an ESE to promote adjacent splice site use. In vitro experiments demonstrated that SF2/ASF can bind the PRE [60, 61]. However, overexpression of SF2/ASF failed to promote use of the splice site adjacent to the PRE but rather increased the level of unspliced viral RNA and reduced the quantity of eRev RNA. In parallel experiments, hnRNP A1 overexpression failed to alter viral RNA splicing patterns [58].

In contrast to HIV-1, there is only a limited understanding of splicing control in human T cell leukemia virus type 1 (HTLV-1). Like HIV-1, HTLV-1 produces several factors, in addition to the structural proteins, by alternative splicing. Little is known of the cis-acting elements controlling splice site use but evidence of regulation is provided by the marked differences in abundance of the various spliced RNA isoforms in different infected cell lines [62, 63]. Overexpression of SF2/ASF or hnRNP A1 alters HTLV-1 RNA splicing patterns [63] and loss of hnRNP A1 expression leads to an accumulation of unspliced viral RNA and increased virus production [64]. Although the effect of hnRNP A1 depletion could be attributed to affects on splicing, the data could also be explained if hnRNP A1 inhibits viral RNA transport to the cytoplasm (putatively by inhibiting binding of the HTLV-1 Rev-like factor, Rex, to the viral RNA) [65].

Several cis-acting elements that affect RNA stability and processing in Moloney murine leukemia virus (MoMLV) have also been identified. Analysis of sequences adjacent to the 3'ss revealed several elements that control splicing [66]. Deletion of exon sequences downstream of the 3'ss resulted in a total loss of spliced viral RNA, suggesting that the region may contain an ESE as seen for EIAV, HIV-1 and ASV. In contrast, deletion of 140 nt immediately upstream of the bpt sequence resulted in a marked elevation in the spliced/unspliced viral RNA ratio, consistent with the presence of a splicing silencer. However, the region surrounding the 3'ss is not the only one that influences splicing. Another element located within the CA region is required for accumulation of spliced viral RNA [67, 68]. This element contrasts with the NRS of ASV as the MoMLV element would appear to be a stimulator of viral RNA splicing. Studies on the Akt strain of MLV identified a region downstream of the 5'ss that modulates splicing efficiency. In the course of examining the contribution of various sequence elements to viral RNA dimer initiation, Aagaard et al. [69] noted that deletion of a stem loop structure (DIS-1) immediately 3' of the 5'ss resulted in a 5–10 fold increase in the level of spliced RNA. Given its close proximity to the 5'ss, the secondary structure of DIS-1 may block base pairing of U1 snRNA to the 5' ss.

In summary, it would appear that retroviruses have used a common set of tools (suboptimal 3'ss, splicing enhancers, splicing silencers) to regulate the extent of viral RNA processing and achieve a balanced level of unspliced and spliced RNAs compatible with virus replication. The use of cellular factors (SR proteins, hnRNP proteins) to regulate splicing suggests that the extent of RNA processing and hence, the capacity of the virus to replicate, is also dictated by the required mix of host cell splicing regulatory factors and determine the host range of the virus. Supporting this conclusion are observations from studies using MLV and HIV-1 based vectors. Lee et al. [70] noted that virus titers from a MLV-based vector varied significantly between various cell lines and showed that at least part of the problem resided in marked differences in viral RNA processing. Since the vector used was constant, the variation seen is likely due to different levels of host splicing factors. A similar phenomenon may also partially explain the inability of murine cells to support HIV-1 replication. Zheng et al. [71] observed excessive splicing of HIV-1 RNA upon introduction of provirus into murine cells. This problem could be alleviated by expression of the human p32 protein, which binds to and likely sequesters SF2/ASF. By reducing the availability of SF2/ASF in this manner, the extent of HIV-1 RNA splicing is reduced, permitting accumulation of genomic RNA. These findings highlight the vulnerability of retroviruses to modulation of host factor regulating RNA processing and raise the possibility of therapeutic intervention at this level.

Polyadenylation

Polyadenylation plays a key role in the life of an mRNA, regulating its transport, translation and turnover. Controlling where in the retroviral genome polyadenylation occurs is critical for replication. For several retroviruses (i.e. HTLV-1, HTLV-2, bovine leukemia virus, RSV, murine leukemia virus), choice of polyadenylation site use is straightforward since the major signal for the reaction (AAUAAA) occurs only once in the transcript. For other retroviruses (i.e. HIV-1, equine infectious anemia, Moloney murine leukemia virus), the situation is rendered more complex by the duplication of the polyadenylation signals (AAUAAA and the 3' G/U-rich sequence) at the 5' and 3' ends of the transcript. Successful replication has necessitated that these viruses evolve mechanisms to suppress recognition of the first polyadenylation signals. Although research has shown that sequences within U3 (present only at the 3' end of the retroviral RNA) can enhance use of the downstream polyadenylation signal, this finding does not readily explain its almost exclusive use.

Regulating HIV-1 RNA polyadenylation

In the case of HIV-1, it was initially believed that the proximity of the first polyadenylation site to the start site of transcription reduced its recognition by the host polyadenylation machinery, possibly as a result of a secondary structure that masks the AAUAAA polyadenylation signal [72, 73]. However, subsequent work has provided an alternative explanation. Inactivation of the first 5'ss (SD1) dramatically increased use of the promoter proximal polyadenylation signal [74–76]. Subsequent experiments determined that recruitment of U1 snRNP to SD1 acts to suppress use of this polyadenylation site, possibly through an interaction between the U1 70 K protein of U1 snRNP and components of the polyadenylation machinery, in particular polyA polymerase [75, 77].

In addition to the cis-acting signals that control use of the first polyadenylation site, ESE3 and ESS3 play a role in modulating use of the second polyadenylation signal. Deletion of ESE3 not only results in decreased use of SA7 but also an inhibition of Rev-dependent viral gene expression [40, 78] that correlated with loss of polyadenylation of the incompletely spliced viral RNA [79]. Polyadenylation of viral RNA could be restored by the deletion of ESS3 and ESE3, indicating that these two elements not only play antagonistic roles in the recognition of SA7 but also in the 3' end processing of viral RNA [79].

Cellular and viral proteins have been implicated in regulating HIV-1 RNA polyadenylation. Experiments with Sam68, a member of the STAR family of RNA binding proteins, revealed that its overexpression dramatically enhanced Rev function [80–83], which correlated with the ability to stimulate 3' end processing of incompletely spliced viral RNA [79]. With the demonstration that Sam68 is essential for both Rev-induced viral gene expression and HIV-1 replication, it would appear that it might play a pivotal role in the processing of the incompletely spliced viral RNAs that renders them competent for transport to the cytoplasm and subsequent translation [84]. Interestingly, the virus itself also modulates the cell's polyadenylation machinery. Vpr-expressing viruses induce a dephosphorylation of poly A polymerase, the enzyme responsible for addition of the poly A tail following cleavage, an alteration that leads to increased activity [85]. While Vpr expression does lead to a modest increase in viral RNA levels, this is not achieved through changes in either viral RNA stability or poly A tail length. However, it remains to be determined whether Vpr might affect the initial processing of the viral transcripts in the nucleus. HIV-1 Tat may also impact on viral RNA polyadenylation through its ability to increase expression of the 73 kDa component of the cleavage and polyadenylation specificity factor (CPSF), a key factor in 3'end processing [86].

MLV: different virus, different solution

Although both HIV-1 and Moloney murine leukemia (MoMLV) virus share the same problem of suppressing use of the 5' proximal polyadenylation signal, they have evolved different mechanisms to solve the problem. In contrast to HIV-1, mutation of the 5'ss in MoMLV has little effect on the use of the first polyadenylation signal. Rather than regulating use of the signal, MoMLV appears to have adopted the use of inefficient signals, resulting in a significant proportion of viral transcripts failing to use either of the two viral encoded polyadenylation signals and the RNA terminates in the adjacent cellular sequences [87].

RSV, the NRS, and 3'-end formation

In the avian RSV, the NRS appears to play an important role in modulating polyadenylation efficiency. Avian retroviral 3'-end formation is inherently inefficient with ~15% of RNA representing read-through transcripts where poly(A) addition occurs at downstream cellular sites [88]. Miller and Stoltzfus [89] showed that deletions encompassing the NRS increased the level of read-through transcripts and proposed that the deleted sequence(s) bind factors that stabilize the poly(A) machinery to allow more efficient polyadenylation. The NRS appears to be the relevant element since specific mutations or deletions within this region also result in 3'-end formation deficiencies [8, 90]. It was proposed that the stalled splicing complex between the NRS and viral 3'ss serves the same function as the splicing process [90]. This model is consistent with observations that NRS mutations induce transcriptional read-through and splicing into the cellular myb gene in chickens, which results in short-latency lymphomas [91]. It will be important in the future to determine the mechanism by which the NRS boosts polyadenylation of genomic RNA.

Nuclear export of incompletely spliced RNA

Once the challenges of manipulating the splicing apparatus to preserve pools of unspliced RNA have been met, retroviruses face the task of exporting these molecules in a cell that normally restricts unspliced RNA to the nucleus. This could involve overcoming nuclear retention signals and/or recruiting export factors that otherwise would have little attraction for genome-length viral RNA. Export of bulk mRNA is thought to be facilitated by the recruitment of the general export factor TAP/NXF1:p15 to the RNA through different adaptor proteins, including REF/Aly and SR proteins [92]. Adaptor loading onto RNA occurs either through an interaction with a mark deposited upon intron removal, such the exon junction complex (EJC), or by direct binding to elements within the RNA. How then do unspliced retroviral RNAs, which don't benefit from the deposition of an EJC, get efficiently exported? As with splicing and polyadenylation, different viruses have evolved distinct mechanisms to export unspliced RNA. In the case of HTLV-1/2 and lentiviruses such as HIV-1, the virus encodes an accessory protein that targets unspliced RNA to an export pathway distinct from that used by most cellular mRNA. In contrast, simple retroviruses like Mason Pfizer monkey virus (MPMV) harbor cis-elements that bind host cell export factors directly, independent of splicing. Moreover, some of these proteins are multifunctional, acting early in splicing regulation and later in RNA trafficking and perhaps viral RNA encapsidation.

Control of HIV-1 RNA export out of the nucleus

Early mutagenesis studies of HIV-1 revealed that loss of Rev expression resulted in a complete loss of viral structural protein expression without significantly affecting levels of the various viral RNAs. Subsequent fractionation studies determined that the absence of HIV-1 structural protein production upon inactivation of the Rev reading frame was due to sequestration of the unspliced 9 kb and singly spliced 4 kb viral RNAs in the nucleus [93–96]. Only the fully processed 2 kb viral mRNAs accumulate in the cytoplasm in the absence of Rev. The basis for the nuclear retention of the 9 kb and 4 kb HIV-1 RNAs remains poorly understood, with some groups attributing it to partial assembly of spliceosomes on the RNA [97], while others have identified cis-acting repressive (CRS) or instability (INS) sequences within the Gag, Pol and Env reading frames that are able to confer Rev-dependency in heterologous contexts [98–107]. As these inhibitory sequences are removed by splicing in the generation of the multiple spliced, 2 kb RNAs, no impediment exists for the transport of these viral RNAs via the general mRNA export pathway of the cell. Additional mutations also demonstrated the requirement of a 240 nt sequence (designated the Rev-responsive element, RRE) within the env reading frame for Rev function [95, 96]. The RRE serves as a point of interaction of viral RNA with the Rev protein.

Intense investigation of HIV-1 Rev function has resulted in it being one of the most thoroughly characterized export systems and readers are referred to more extensive reviews on its function that are briefly summarized here [95, 96]. Multiple domains are required for Rev to function. Within the amino terminal portion of Rev is an arginine-rich stretch between a.a. 35–50 that comprises a nuclear/nucleolar localization signal (NLS/NoLS) and forms an alpha helix able to bind in the major groove of the primary Rev-binding site of RRE RNA. Within the carboxyl terminal portion is a leucine-rich sequence between a.a. 73 and 84 that forms the nuclear export signal (NES). Despite steady-state accumulation in the nucleolus, the presence of both an NLS and NES within Rev results in the protein constantly moving between the nucleus and cytoplasm. Nuclear import is mediated by binding of the arginine-rich region to the transport mediator importin β and nuclear export is achieved through binding of Crm1/Exportin-1 to the leucine-rich NES of Rev in a Ran/GTP dependent manner. This ternary complex (Rev/Crm1/RanGTP) then interacts with the FG-repeats of nucleoporins and the complex moves through the nuclear pore. Once within the cytoplasm, the RanGTP within the complex is hydrolyzed to RanGDP by RanGAP and the ternary complex disassembles.

Although it is possible that Rev interacts with all RRE-containing HIV-1 RNAs in the nucleus (the 9 and 4 kb class of RNAs), studies have indicated that several parameters dictate which RNA will be exported to the cytoplasm. First was the demonstration that Rev function was dependent upon the continued transcription of the target RNA despite the presence of significant levels of RRE-containing RNA in the nucleus [108]. This finding suggests that Rev must act before the viral RNA either becomes fully spliced or is committed to retention in the nucleus. Second was the observation that Rev-induced export required 3' end processing of the RNA as only polyadenylated viral RNAs are transported to the cytoplasm [79, 109]. Therefore, while Rev is able to bypass the cellular mechanisms that prevent export of incompletely spliced RNAs from nucleus, the affected RNAs must meet a limited set of criteria (5' cap, 3' poly A tail) to be exported. The requirement for 3' end processing for Rev-mediated export may provide a partial explanation for the need for continued synthesis of the target RNA. Recent studies have established that a tight coupling exists between the various processing steps leading to mature mRNAs, suggesting that once 3' end formation occurs it would stimulate the removal of the upstream intron [110–112]. Therefore Rev may need to act within the brief time frame between 3' polyadenylation and subsequent splicing of the RNA to induce export of the unspliced and partially spliced viral RNAs. The population of incompletely spliced HIV-1 RNAs that fail to become polyadenylated are likely retained in the nucleus and degraded.

Once the Rev/Crm-1/RanGTP complex assembles on the appropriate RNA, its journey from the site of synthesis to the cytoplasm begins. Most of the details of this process remain unclear but recent experiments have begun to identify host factors that play pivotal roles in the process. At least two members of the DEAD box RNA helicase family, DDX3 and DDX1, play essential roles in mediating Rev-dependent RNA export [113, 114]. Depletion of either protein is associated with a marked reduction in Rev activity [113, 114]. DDX1 also interacts with Rev via the N-terminal domain, suggesting a role in initial complex assembly [114]. For DDX3, its interaction with CRM-1 and localization to the outer nuclear membrane suggests that it might act to facilitate the translocation of the Rev-RNA complex through the nuclear pore [113]. Once on the cytoplasmic face of the nuclear membrane, another host factor, hRIP, appears to be required for release of the viral RNA into the cytoplasm [115, 116] as depletion of the protein results in accumulation of viral RNA on the cytoplasmic face of the nucleus. A similar perinuclear accumulation of viral RNA is also observed upon overexpression of a C-terminal deletion mutant of Sam68, designated Sam68ΔC [83]. However, in this instance subsequent studies (Marsh and A.C., unpublished) indicate that this factor acts at a later step in the cytoplasmic metabolism of the viral RNA.

CTE pathway

Although the study of the HIV-1/lentivirus systems clearly demonstrated a role for Rev-like proteins as adaptors to facilitate the export of viral RNAs to the cytoplasm, parallel work indicated that other viruses evolved alternative solutions to the export problem. An element at the 3' end of the MPMV designated the constitutive transport element (CTE) was shown to support Rev-independent HIV structural protein expression [117–119]. An element with similar activity was found in simian retrovirus type I [120]. The CTE is able to competitively inhibit cellular mRNA export (unlike Rev or the RRE) and interacts with the host export factor NXF1 [121–123]. Thus, in contrast to HIV-1 where Rev serves as an adaptor to access the export pathway, direct binding of NXF1 to the CTE bypasses the requirement for a virally-encoded protein.

DR1/DR2 elements of avian retroviruses

Identification of the CTE export element in MPMV implied that similar export elements would be found in other simple retroviruses. One such element required for cytoplasmic accumulation of Pr-C RSV unspliced RNA was localized to the direct repeat region downstream of the src gene (DR2) [124]. A second repeat (DR1) located upstream of src shows similar activity, and at least one DR is required for RSV replication. The DR sequence is conserved between avian retroviruses and a similar activity was ascribed to the DR element in RAV-2 ALV [125]. Despite the clear export activity of the DR elements in reporter assays, unambiguous demonstration of an export role in RSV is complicated by the finding that DR deletions have effects apart from export, including destabilization of unspliced RNA and defective particle assembly [124, 126, 127]. While the DR elements harbor export activity, they are not absolutely required for export since the results from Simpson et al. [126] indicate that some unspliced RNA is exported and translated even in their absence.

As discussed above, HIV-1 Rev serves as an adaptor to target HIV RNA to the CRM1 export pathway, whereas the MPMV CTE directly binds NXF1 for export via the mRNA route. There is no obvious sequence similarity between the MPMV CTE and the avian DR elements, but it is perhaps reasonable to speculate that other simple retroviruses evolved to exploit the NXF1 pathway for export. Two studies took advantage of reagents to block the CRM1 pathway and found that, like the MPMV CTE, export of DR reporter RNAs was unaffected under conditions that blocked Rev export [125, 128]. Thus, the CRM1 pathway is not required for ALV RNA export. Surprisingly, NXF1 binding to the ASV DR elements has yet to be demonstrated, implying that avian retroviral unspliced RNA export exploits yet a different pathway from HIV-1 and MPMV. However, the possibility that the avian DR elements bind a cellular adaptor molecule that functions similarly to NXF1, or that interacts with NXF1, has not been eliminated.

The importance of the road traveled to the cytoplasm

It is well established that the nuclear history of an mRNA can influence its fate in the cytoplasm. This property can be attributed to the nature of the mRNP that assembles on the RNA. One well-studied example is the affect that mRNA splicing has on mRNP composition through the deposition of the exon junction complex (EJC) and its consequences to downstream events such as export, translation, and decay [129]. The possibility that cis elements that direct retroviral RNAs to one export pathway or another might influence downstream cytoplasmic events was realized with the observation that unspliced RSV RNAs that lack DRs produce readily detectable amounts of Gag protein but are defective in particle assembly [126, 127]. These investigators hypothesized that either the RNA export defect rendered Gag synthesis below a threshold level required for assembly or more intriguingly, that the DR contains an element distinct from that responsible for CTE activity and directs RNA to a cytoplasmic location that is conducive to production of assembly-competent Gag protein.

Mammalian cells are nonpermissive for ALV infection due to defects in RNA processing, RNA export, Gag cleavage and particle assembly [124, 126, 130, 131]. These observations are similar to those reported for ΔDR viruses in avian cells and suggest that the RNP exported in mammalian cells fails to deliver genome-length RNA to a cytoplasmic location where translated Gag can assemble particles. [124, 126]. This idea is supported by work demonstrating that ALV particles can be formed in mammalian cells when the RRE is provided in cis and Rev in trans, i.e., when RNA is exported via the CRM1 pathway [106]. This result seems at odds with the lack of an effect of CRM1 inhibitors in avian cells, but it is possible that productive export occurs by more than one pathway, as is true for HIV-1 (see below).

A recent report by Swanson et al. demonstrated a similar link between HIV-1 unspliced RNA export and Gag assembly [132]. Gag protein can be produced in murine cells but is not processed or assembled into virions, which is one reason that murine cells are nonpermissive for HIV-1 replication. These investigators demonstrated that rerouting unspliced RNA export from the Rev/RRE pathway to the CTE pathway restored efficient virion production. This correlated with a redistribution of Gag from diffuse cytoplasmic localization when RNA export was Rev/RRE-dependent to plasma membrane association when the CTE route was used. Particle assembly occurs in human cells regardless of the export pathway used by the Gag RNA. Thus, as with ALV, productive HIV-1 RNA export may produce some type of RNP 'mark' that influences cytoplasmic RNA localization and the ability of the encoded protein to reach assembly sites. Part of this mark could lie in the association of hnRNP proteins on the unspliced RNAs at a particular step of the viral gene expression phase. Bériault et al. showed that disruption of hnRNP A2 binding to its cognate cis sequence (the hnRNP A2 response element or A2RE) also affected the cellular distribution of Gag and the auxiliary protein Vpr, but only at a late step that coincided with a block in unspliced HIV-1 RNA export from the nucleus [133]. It will be important to determine the composition of the RNPs that are and are not competent to direct productive Gag synthesis. This is clearly an area that deserves more research as it likely represents an interface between nuclear events and the formation/function of possible intracellular transport granules.

Intracytoplasmic trafficking of retroviral RNA

Once delivered into the cytoplasm, two fates exist for the unspliced, genomic viral RNA: translation to produce the structural proteins and selection for encapsidation into the forming virions. The majority of unspliced retroviral RNA is not captured for encapsidation but serves other roles in generating viral structural proteins and enzymes or as a cofactor for assembly ([134–137] and reviewed in [138]). However, genomic viral RNA that is translated in the cytoplasm must transition in some fashion to sites of virus assembly to become encapsidated. The first evidence suggesting a specific location for genomic RNA selection for encapsidation has recently come to light in work from Andrew Lever's group. Using fluorescence resonance energy transfer (FRET), they were able to monitor the interaction of Gag with unspliced viral RNA. Unexpectedly, the unspliced HIV-1 RNA was found to be captured by Gag at a site at or adjacent to the centriole, near the nuclear membrane [139]. The signal was dependent upon psi-containing viral RNA. This was an infrequent event, consistent with earlier reports indicating that the vast majority of the unspliced viral RNA is not selected for encapsidation but translated or used as a cofactor for assembly [140, 141]. The centriole region has also been identified as an assembly site for type D assembling retroviruses such as MPMV [142, 143]. Image analysis reveals translating polyribosomes and co-assembly of capsids in this region but it remains unclear if assembly of retroviral type C HIV-1 capsids are also initiated in this region. This binding event could represent the first step in the formation of an RNP transport complex (see below) to sites of capsid formation. FRET has also been used successfully to identify cellular regions at which Gag-Gag homo-oligomerization in membranes occurs during viral assembly [144, 145] and these types of techniques will help decipher some of the molecular interactions during viral replication.

Deciphering the relationship between the centriole, Gag capture and encapsidation into virions, and the process that directs viral genomic RNA (and possibly other viral mRNAs) to the sites of assembly remains a considerable challenge requiring targeting signals in the viral RNA, viral proteins, and/or a host cell targeting machinery [146]. While RNA can move intracellularly by a variety of mechanisms (Brownian motion, active transport) [147], clues about this process for retroviruses were provided by pioneering RNA trafficking studies in which vesicular trafficking pathways were shown to deliver viral components to assembly sites. Part of this newly described retroviral RNA trafficking pathway relies on the recruitment of genomic RNA from a cytoplasmic pool onto vesicles.

Retroviral RNA trafficking on cellular vesicles in the cytoplasm

The movement of MLV RNA to sites of virion assembly was investigated by monitoring MLV genomic RNA movement in live cells using a bacteriophage MS2 tethering system. In this study [148], it was shown that genomic RNA traffics on recycling endosomal vesicles. Time-lapse fluorescence video microscopy showed a directed and linear trafficking pathway that was dependent upon the integrity of the microtubules [148]. Transport required the psi RNA packaging signal within the affected RNA and an intact NC domain in Gag, consistent with their demonstrated requirements for viral RNA packaging. The vesicles were comprised of both endosomal and lysosomal vesicles as evidenced by co-trafficking of the labeled RNA on transferrin- and lysotracker-positive vesicles in cells. Results of experiments in which monensin was used, a drug that prevents acidification of endosomal and organellar compartments, indicated that trafficking was achieved on vesicles that likely emanate from a steady-state endosomal compartment and not rapidly recycling vesicles that contain Rab11. Gag protein is recruited from late endosomal/lysosomal compartments to these endosomal membranes by the viral glycoprotein, Env, demonstrating important contributions of Env to this process. Not only could cellular RNAs replace viral RNA on vesicles in their system when psi was mutated, but some evidence suggests that the psi RNA sequence that is comprised of four stem loops in MLV, harbours an endosomal trafficking signal [149]. Other studies suggested a similar role for recycling membrane compartments and the expression of Env in the assembly of MPMV virus particles [143], although RNA trafficking was not examined. Despite the classical differences in the assembly pathways used by the two viruses (MLV is a type C virus for which capsid assembly and morphogenesis occur at the plasma membrane while MPMV capsids assemble intracellularly at the centriole [143]) the similarities in intracellular trafficking pathways that rely on recycling membrane compartments and the requirement for Env in the assembly of virus particles warrants further investigation [143]. Our own results add to this story with the demonstration that Env expression can dramatically alter the distribution of HIV-1 genomic RNA in HeLa cells (K. Lévesque, M. Halvorsen & A.J.M., unpublished). Basyuk and colleague's work supports earlier evidence that several retroviral Gags interact with kinesin motor proteins to enable trafficking along microtubules. The significance of these observations is yet to be deciphered but might put Gag as the key component of these trafficking complexes. Both Gag and RNA were visualized on the outside of translocating vesicles in the absence of MLV capsids, suggesting that part of this trafficking pathway is preceded by the formation of a cytosolic RNP complex [148]. While it in not known where the recruitment of MLV Gag and RNA occur, some insight was provided by the work of Poole's et al. in which HIV-1 RNA capture by Gag was shown to occur at the centriole [139]. MPMV RNA appears to be cotranslationally assembled in this cellular region, suggesting that this may be a point where the viral RNA transitions from a free RNP particle into a membrane-bound complex en route to the plasma membrane. It remains to be determined whether these relationships hold true for all retroviruses.

The RNA trafficking granule takes shape

The directed movement of viral RNA within the cytoplasm relies on its interaction with multiple host proteins generating an RNA transport granule (RTG). The concept of an RTG derives from the studies in neuronal cells, which have both specialized functions and extended morphologies [150–152]. RTGs were shown to contain translational components such as transfer-RNA synthetases, EF1α, ribosomal RNAs, and molecular motor proteins such as dynein and kinesin [153] (Table 1). Although characterized in specialized neuronal cells, the RTG likely exists in some form in most other cell types, such as in fibroblasts, T cells and epithelial cells [154, 155], but the morphologies of some cell types make it difficult to study RNA trafficking events (e.g., T cells). The RTG is indeed assembled in oligodendrocytes and each granule can contain multiple copies of HIV-1 mRNAs [151].

Table 1 Links between components of RNA trafficking granules and retroviral replication

Published reports provide ample evidence that components of the RTG play roles in multiple steps of retroviral replication including transcription, RNA splicing, nucleocytoplasmic transport, translation as well as genomic RNA encapsidation (see Table 1). The effect of these factors on RNA transport within neuronal cells and the identified roles for many of these proteins in retrovirus replication highlight a potential functional relationship between the RTG machinery and retroviral replication. Retroviruses might co-opt components of this cellular machinery to ensure both the correct trafficking and localization of the retroviral RNA for presentation to the translation machinery and at sites of viral assembly for encapsidation into virions [146]. The proteins found in the RTG are presented in Table 1 and a subset are discussed below.

RNA helicases: RHA, DDX1 & DDX3

As described above, nuclear export of retroviral RNA involves several cellular RNA helicases. Recent observations have identified roles for RNA helicase A (RHA) and two DEAD-box proteins, DDX1 and DDX3 in the nuclear export of retroviral RNAs [113, 156–158]. The functions of the latter proteins have been reviewed earlier [159] and are described in a previous section of this review. These RNA helicase proteins appear to act at different stages of retroviral gene expression. RHA depletion by siRNA decreases translation of HIV-1 gag-pol mRNA, perhaps by disrupting the remodeling of RNA-RNA and RNA-protein interactions that are required for usage of unspliced transcripts by the translation apparatus (K. Boris-Lawrie, unpublished results). The recent identification of these proteins in RTGs in the cytoplasm [160], albeit in specialized neuronal cells, leads to the idea that they may represent part of the "protein mark" that initially tags retroviral RNAs in the nucleus and remains associated during nuclear export and subsequent RTG assembly in the cytoplasm. These findings suggest a possible mechanism by which the nuclear history of the retroviral RNA and the associated proteins might dictate gene expression patterns in the cytoplasm analogous to the role of the exon junction complex [161, 162].

hnRNP A2

hnRNPs in general are believed to function at many post-transcriptional levels (reviewed in [163]). Particular attention has been placed on their role in the retroviral RNA processing and nuclear export [39, 64, 164, 165]. Some of the evidence that particular hnRNPs could play other -yet related- roles in retroviral replication such as RNA trafficking are outlined here. Two 21 nucleotide sequences in HIV-1 RNA were shown to bear striking homology to the myelin basic protein (MBP) A2RE located in the 3' untranslated region of the corresponding mRNA. The MBP A2RE is required for RNA trafficking to the extremity of dendrites of murine oligodendrocytes [150, 166]. Injected HIV-1 and HIV-2 RNAs are also efficiently trafficked in these cells with requirements identical to those described for MBP mRNA: an intact A2RE, hnRNP A2 expression, microtubules and kinesin ([151] & A.J.M, E. Barbarese, É.A. Cohen, J. Carson, unpublished results). By sequence comparison, similar cis-acting elements can also be found in HTLV-1 and HIV-2 RNAs but only the element in HIV-2 vpr RNA has been shown to be active in RNA trafficking [151]. Subsequent studies using proviral HIV-1 constructs confirmed these results and demonstrated that hnRNP A2 association with A2RE is critical for late viral gene expression and genomic RNA encapsidation [133]. Kinesin motor proteins are critical for hnRNP A2-mediated RNA trafficking in oligodendrocytes and hnRNP A2 was recently shown to interact with microtubule adaptor proteins [167] and is present in RNA transport granules [160]. This observation provides a physical link to the cytoskeleton on which organelles, vesicles, and RTGs are transported. The interactions between retroviral Gag proteins and the kinesin motor protein, KIF-4, also suggest that large multi-protein complexes are translocated on the microtubule-based cytoskeleton in virus-expressing cells [168, 169]. These examples provide evidence that viral RNP translocation is dictated by the activities of a variety of viral and cellular proteins.

Staufen1

While many of the proteins within the RTG have functions in several cellular compartments, Staufen1 works mainly in the cytoplasm. Staufen1 represents a bona fide RNA trafficking protein whose function is conserved from lower to higher eukaryotes. It plays a critical role in the localization of several mRNAs in Drosophila, likely via its direct interaction with target RNAs or in the context of RNP complexes [170, 171]. Several lines of evidence implicate Staufen1 in regulating retroviral genomic RNA encapsidation. First, Staufen1 associates with precursor Gag protein and not the mature Gag proteins, and preferentially co-precipitates with the genomic viral RNA but not spliced forms [172]. Second, Staufen1 is encapsidated in virions in stoichiometry to the number of genomic RNA molecules present. Its identification in other retrovirus particles suggests that Staufen1 may have a more general role in the selection of genomic RNA for retroviral encapsidation. Third, Drosophia Staufen preferentially binds RNA homodimers [173], the form viral genomic RNA takes during virion morphogenesis and within retrovirus particles [174]. Fourth, Staufen1 knockdown by siRNA results in a dramatic decrease in infectivity and virus production [172]. Fifth and more importantly, modulation of intracellular levels of Staufen1 directly impacts on genomic RNA encapsidation [172, 175]; L. Abrahamyan, J.-F. Clément & A.J.M., unpublished results). These findings suggest that Staufen1 may tag the genomic RNA for encapsidation in the cytoplasm and be concomitantly recruited into virus particles. Its function in this process needs to be unequivocally proven, however published data suggest that a molecular switch could be at play that promotes the selection of two copies of genomic RNA per virion [172]. Likely, the activity of Staufen1 is but one of several factors needed for the "genomic RNA selection" process. It will be important to evaluate how the functions of this and other RNA trafficking proteins are interrelated along the retroviral RNA targeting pathway from the nucleus to progeny virions.

Conclusion

Several functional and physical links between the transcription, RNA processing, nucleocytoplasmic and cytoplasmic transport machineries as well as the metabolic machineries for RNA in the cytoplasm are made by RNA-binding proteins, some of which mark the RNA before their exit from the nucleus and modulate RNA fate in the cytoplasm. The journey of retroviral RNA from the nucleus to its ultimate destination in the capsid is likely to be characterized by similar phenomena, yet up until now the details of these events have been scant. Many details of how the RNA processing machinery is manipulated to produce the appropriate spectrum of mRNAs and to allow genome-length RNA to escape splicing have been revealed in the last few years and illuminated potential targets for arresting virus replication. Recent work has already exploited our understanding of retroviral RNA processing to develop small molecule inhibitors of HIV-1 RNA splicing [176]. Another critical step in replication is the nuclear export of genome-length RNA and efforts are being made towards gaining a better understanding of the processes that remodel nuclear RNPs and determine the fates of viral RNA within the cytoplasm. A number of factors with roles in cytoplasmic RNA transport have been found associated with retroviral mRNA. A more detailed understanding of the formation and function of retroviral RNA trafficking granules and their components may provide new insights into targets for rationale drug design, an idea that recently has met with some success [177].

References

  1. Fu XD, Katz RA, Skalka AM, Maniatis T: The role of branchpoint and 3' exon sequences in the control of balanced splicing of avian retrovirus RNA. Genes & Development. 1991, 211-220.

    Google Scholar 

  2. Katz RA, Skalka AM: Control of retroviral RNA splicing through maintenance of suboptimal processing signals. Molecular & Cellular Biology. 1990, 10 (2): 696-704.

    CAS  Google Scholar 

  3. Zhang L, Stoltzfus CM: A suboptimal src 3' splice site is necessary for efficient replication of Rous sarcoma virus. Virology. 1995, 206 (2): 1099-1107. 10.1006/viro.1995.1033.

    CAS  PubMed  Google Scholar 

  4. Amendt BA, Simpson SB, Stoltzfus CM: Inhibition of RNA splicing at the Rous sarcoma virus src 3' splice site is mediated by an interaction between a negative cis element and a chicken embryo fibroblast nuclear factor. J Virol. 1995, 69 (8): 5068-5076.

    PubMed Central  CAS  PubMed  Google Scholar 

  5. McNally MT, Beemon K: Intronic sequences and 3' splice sites control Rous sarcoma virus RNA splicing. Journal of Virology. 1992, 66 (1): 6-11.

    PubMed Central  CAS  PubMed  Google Scholar 

  6. Arrigo S, Beemon K: Regulation of Rous sarcoma virus RNA splicing and stability. Molecular & Cellular Biology. 1988, 8: 4858-4867.

    CAS  Google Scholar 

  7. Stoltzfus CM, Fogarty SJ: Multiple regions in the Rous sarcoma virus src gene intron act in cis to affect the accumulation of unspliced RNA. Journal of Virology. 1989, 63 (4): 1669-1676.

    PubMed Central  CAS  PubMed  Google Scholar 

  8. O'Sullivan CT, Polony TS, Paca RE, Beemon KL: Rous sarcoma virus negative regulator of splicing selectively suppresses SRC mRNA splicing and promotes polyadenylation. Virology. 2002, 302 (2): 405-412. 10.1006/viro.2002.1616.

    PubMed  Google Scholar 

  9. McNally MT, Gontarek RR, Beemon K: Characterization of Rous sarcoma virus intronic sequences that negatively regulate splicing. Virology. 1991, 185 (1): 99-108. 10.1016/0042-6822(91)90758-4.

    CAS  PubMed  Google Scholar 

  10. Hibbert CS, Gontarek RR, Beemon KL: The role of overlapping U1 and U11 5' splice site sequences in a negative regulator of splicing. RNA. 1999, 5 (3): 333-343. 10.1017/S1355838299981347.

    PubMed Central  CAS  PubMed  Google Scholar 

  11. McNally LM, McNally MT: U1 small nuclear ribonucleoprotein and splicing inhibition by the rous sarcoma virus negative regulator of splicing element. Journal of Virology. 1999, 73 (3): 2385-2393.

    PubMed Central  CAS  PubMed  Google Scholar 

  12. Paca RE, Hibbert CS, O'Sullivan CT, Beemon KL: Retroviral splicing suppressor requires three nonconsensus uridines in a 5' splice site-like sequence. Journal of Virology. 2001, 75 (16): 7763-7768. 10.1128/JVI.75.16.7763-7768.2001.

    PubMed Central  CAS  PubMed  Google Scholar 

  13. Patel AA, Steitz JA: Splicing double: insights from the second spliceosome. Nat Rev Mol Cell Biol. 2003, 4 (12): 960-970. 10.1038/nrm1259.

    CAS  PubMed  Google Scholar 

  14. Gontarek RR, McNally MT, Beemon K: Mutation of an RSV intronic element abolishes both U11/U12 snRNP binding and negative regulation of splicing. Genes & Development. 1993, 7 (10): 1926-1936.

    CAS  Google Scholar 

  15. Fogel BL, McNally MT: A cellular protein, hnRNP H, binds to the negative regulator of splicing element from rous sarcoma virus. Journal of Biological Chemistry. 2000, 275 (41): 32371-32378. 10.1074/jbc.M005000200.

    CAS  PubMed  Google Scholar 

  16. McNally LM, McNally MT: SR protein splicing factors interact with the Rous sarcoma virus negative regulator of splicing element. Journal of Virology. 1996, 70 (2): 1163-1172.

    PubMed Central  CAS  PubMed  Google Scholar 

  17. Cook CR, McNally MT: SR protein and snRNP requirements for assembly of the Rous sarcoma virus negative regulator of splicing complex in vitro. Virology. 1998, 242: 211-220. 10.1006/viro.1997.8983.

    CAS  PubMed  Google Scholar 

  18. Cook CR, McNally MT: Characterization of an RNP complex that assembles on the Rous sarcoma virus negative regulator of splicing element. Nucleic Acids Research. 1996, 24 (24): 4962-4968. 10.1093/nar/24.24.4962.

    PubMed Central  CAS  PubMed  Google Scholar 

  19. Cook CR, McNally MT: Interaction between the negative regulator of splicing element and a 3' splice site: requirement for U1 small nuclear ribonucleoprotein and the 3' splice site branch point/pyrimidine tract. Journal of Virology. 1999, 73 (3): 2394-2400.

    PubMed Central  CAS  PubMed  Google Scholar 

  20. Montzka KA, Steitz JA: Additional low-abundance human small nuclear ribonucleoproteins: U11, U12, etc. Proc Natl Acad Sci U S A. 1988, 85 (23): 8885-8889.

    PubMed Central  CAS  PubMed  Google Scholar 

  21. McNally LM, Yee L, McNally MT: Two regions promote U11 snRNP binding to a retroviral splicing inhibitor element (NRS). J Biol Chem. 2004, 204: 38201-38208. 10.1074/jbc.M407073200.

    Google Scholar 

  22. McNally LM, Yee L, McNally MT: Heterogeneous nuclear ribonucleoprotein H is required for optimal U11 small nuclear ribonucleoprotein binding to a retroviral RNA-processing control element: implications for U12-dependent RNA splicing. J Biol Chem. 2006, 281 (5): 2478-2488. 10.1074/jbc.M511215200.

    CAS  PubMed  Google Scholar 

  23. Caputi M, Zahler AM: SR proteins and hnRNP H regulate the splicing of the HIV-1 tev-specific exon 6D. Embo J. 2002, 21 (4): 845-855. 10.1093/emboj/21.4.845.

    PubMed Central  CAS  PubMed  Google Scholar 

  24. Han K, Yeo G, An P, Burge CB, Grabowski PJ: A combinatorial code for splicing silencing: UAGG and GGGG motifs. PLoS Biol. 2005, 3 (5): e158-10.1371/journal.pbio.0030158.

    PubMed Central  PubMed  Google Scholar 

  25. Garneau D, Revil T, Fisette JF, Chabot B: Heterogeneous Nuclear Ribonucleoprotein F/H Proteins Modulate the Alternative Splicing of the Apoptotic Mediator Bcl-x. J Biol Chem. 2005, 280 (24): 22641-22650. 10.1074/jbc.M501070200.

    CAS  PubMed  Google Scholar 

  26. Tang H, Kuhen KL, Wong-Staal F: Lentivirus replication and regulation. Annual Review of Genetics. 1999, 33: 133-170. 10.1146/annurev.genet.33.1.133.

    CAS  PubMed  Google Scholar 

  27. Schwartz S, Felber BK, Benko DM, Fenyo EM, Pavlakis GN: Cloning and functional analysis of multiply spliced mRNA species of human immunodeficiency virus type 1. Journal of Virology. 1990, 64: 2519-2529.

    PubMed Central  CAS  PubMed  Google Scholar 

  28. Purcell D, Martin MA: Alternative splicing of human immunodeficiency virus type 1 mRNA modulates viral protein expression, replication, and infectivity. J Virol. 1993, 67: 6365-6378.

    PubMed Central  CAS  PubMed  Google Scholar 

  29. O'Reilly M, McNally M, Beemon K: Two strong 5' splice sites and competing 3' splice sites involved in alternative splicing of human immunodeficiency virus type 1 RNA. Virology. 1995, 213: 373-385. 10.1006/viro.1995.0010.

    PubMed  Google Scholar 

  30. Si Z, Amendt BA, Stoltzfus CM: Splicing efficiency of human immunodeficiency virus type 1 tat RNA is determined by both a suboptimal 3' splice site and a 10 nucleotide exon splicing silencer element located within tat exon 2. Nucleic Acids Research. 1997, 25 (4): 861-867. 10.1093/nar/25.4.861.

    PubMed Central  CAS  PubMed  Google Scholar 

  31. Staffa A, Cochrane A: The tat/rev Intron of Human Immunodeficiency Virus Type 1 Is Inefficiently Spliced because of Suboptimal Signals in the 3' Splice Site. Journal of Virology. 1994, 68: 3071-3079.

    PubMed Central  CAS  PubMed  Google Scholar 

  32. Amendt B, Si Z, Stoltzfus CM: Presence of Exon Splicing Silencers within Human Immunodeficiency Virus Type 1 tat Exon 2 and tat-rev Exon 3: Evidence for Inhibition Mediated by Cellular Factors. Molecular and Cellular Biology. 1995, 15: 4606-4615.

    PubMed Central  CAS  PubMed  Google Scholar 

  33. Amendt BA, Hesslein D, Chang LJ, Stoltzfus CM: Presence of Negative and Positive cis-Acting RNA Splicing Elements within and Flanking the First tat Coding Exon of Human Immunodeficiency Virus Type 1. Molecular and Cellular Biology. 1994, 14: 3960-3970.

    PubMed Central  CAS  PubMed  Google Scholar 

  34. Bilodeau PS, Domsic JK, Mayeda A, Krainer AR, Stoltzfus CM: RNA splicing at human immunodeficiency virus type 1 3' splice site A2 is regulated by binding of hnRNP A/B proteins to an exonic splicing silencer element. Journal of Virology. 2001, 75 (18): 8487-8497. 10.1128/JVI.75.18.8487-8497.2001.

    PubMed Central  CAS  PubMed  Google Scholar 

  35. Jacquenet S, Mereau A, Bilodeau PS, Damier L, Stoltzfus C, Branlant C: A Second Exon Splicing Silencer within the Human Immunodeficiency Virus Type 1 tat Exon 2 Represses Splicing of Tat mRNA and Binds Protein hnRNP H. J Biol Chem. 2001, 276: 40464-40475. 10.1074/jbc.M104070200.

    CAS  PubMed  Google Scholar 

  36. Si ZH, Rauch D, Stoltzfus M: The Exon Splicing Silencer in the Human Immunodeficiency Virus Type 1 Tat Exon 3 Is Bipartite and Acts Early in Spliceosome Assembly. Molecular and Cellular Biology. 1998, 18: 5404-5413.

    PubMed Central  CAS  PubMed  Google Scholar 

  37. Staffa A, Cochrane A: Identification of positive and negative splicing regulatory elements within the terminal tat-rev exon of human immunodeficiency virus type 1. Molecular and Cellular Biology. 1995, 15: 4597-4605.

    PubMed Central  CAS  PubMed  Google Scholar 

  38. Domsic JK, Wang Y, Mayeda A, Krainer AR, Stoltzfus CM: Human immunodeficiency virus type 1 hnRNP A/B-dependent exonic splicing silencer ESSV antagonizes binding of U2AF65 to viral polypyrimidine tracts. Molecular & Cellular Biology. 2003, 23 (23): 8762-8772. 10.1128/MCB.23.23.8762-8772.2003.

    CAS  Google Scholar 

  39. Caputi M, Mayeda A, Krainer AR, Zahler AM: hnRNP A/B proteins are required for inhibition of HIV-1 pre-mRNA splicing. Embo J. 1999, 18 (14): 4060-4067. 10.1093/emboj/18.14.4060.

    PubMed Central  CAS  PubMed  Google Scholar 

  40. Asai K, Platt C, Cochrane A: Control of HIV-1 env RNA splicing and transport: investigating the role of hnRNP A1 in exon splicing silencer(ESS3a) function. Virology. 2003, 314 (1): 229-242. 10.1016/S0042-6822(03)00400-8.

    CAS  PubMed  Google Scholar 

  41. Zhu J, Mayeda A, Krainer A: Exon identity established through differential antagonism between exonic splicing silencer-bound hnRNP A1 and enhancer-bound SR proteins. Molecular Cell. 2001, 8 (6): 1351-1361. 10.1016/S1097-2765(01)00409-9.

    CAS  PubMed  Google Scholar 

  42. Marchand V, Mereau A, Jacquenet S, Thomas D, Mougin A, Gattoni R, Stevenin J, Branlant C: A Janus splicing regulatory element modulates HIV-1 tat and rev mRNA production by coordination of hnRNP A1 cooperative binding. Journal of Molecular Biology. 2002, 323 (4): 629-652. 10.1016/S0022-2836(02)00967-1.

    CAS  PubMed  Google Scholar 

  43. Tange TO, Damgaard CK, Guth S, Valcarcel J, Kjems J: The hnRNP A1 protein regulates HIV-1 tat splicing via a novel intron silencer element. EMBO Journal. 2001, 20 (20): 5748-5758. 10.1093/emboj/20.20.5748.

    PubMed Central  CAS  PubMed  Google Scholar 

  44. Blanchette M, Chabot B: Modulation of exon skipping by high affinity hnRNP A1-binding sites and by intron elements that repress splice site utilization. EMBO Journal. 1999, 18: 1939-1952. 10.1093/emboj/18.7.1939.

    PubMed Central  CAS  PubMed  Google Scholar 

  45. Nasim FU, Hutchison S, Cordeau M, Chabot B: High-affinity hnRNP A1 binding sites and duplex-forming inverted repeats have similar effects on 5' splice site selection in support of a common looping out and repression mechanism. Rna. 2002, 8 (8): 1078-1089. 10.1017/S1355838202024056.

    PubMed Central  CAS  PubMed  Google Scholar 

  46. Caputi M, Freund M, Kammler S, Asang C, Schaal H: A bidirectional SF2/ASF- and SRp40-dependent splicing enhancer regulates human immunodeficiency virus type 1 rev, env, vpu, and nef gene expression. Journal of Virology. 2004, 78 (12): 6517-6526. 10.1128/JVI.78.12.6517-6526.2004.

    PubMed Central  CAS  PubMed  Google Scholar 

  47. Kammler S, Leurs C, Freund M, Krummheuer J, Seidel K, Tange TO, Lund MK, Kjems J, Scheid A, Schaal H: The sequence complementarity between HIV-1 5' splice site SD4 and U1 snRNA determines the steady-state level of an unstable env pre-mRNA. RNA. 2001, 7 (3): 421-434. 10.1017/S1355838201001212.

    PubMed Central  CAS  PubMed  Google Scholar 

  48. Mayeda A, Screaton G, Chandler S, Fu XD, Krainer A: Substrate Specificities of SR Proteins In Constitutive Splicing Are Determined by Their RNA Recognition Motifs and Composite pre-mRNA Exonic Elements. Molecular and Cellular Biology. 1999, 19: 1853-1863.

    PubMed Central  CAS  PubMed  Google Scholar 

  49. Tange TO, Kjems J: SF2/ASF binds to a splicing enhancer in the third HIV-1 tat exon and stimulates U2AF binding independently of the RS domain. Journal of Molecular Biology. 2001, 312 (4): 649-662. 10.1006/jmbi.2001.4971.

    CAS  PubMed  Google Scholar 

  50. Zahler AM, Damgaard CK, Kjems J, Caputi M: SC35 and heterogeneous nuclear ribonucleoprotein A/B proteins bind to a juxtaposed exonic splicing enhancer/exonic splicing silencer element to regulate HIV-1 tat exon 2 splicing. Journal of Biological Chemistry. 2004, 279 (11): 10077-10084. 10.1074/jbc.M312743200.

    CAS  PubMed  Google Scholar 

  51. Manley JL, Tacke R: SR proteins and splicing control. Genes and Development. 1996, 10: 1569-1579.

    CAS  PubMed  Google Scholar 

  52. Zahler AM, Lane WS, Stolk JA, Roth MB: SR proteins: a conserved family of pre-mRNA splicing factors. Genes and Development. 1992, 6: 837-847.

    CAS  PubMed  Google Scholar 

  53. Zahler AM, Neugebauer KM, Lane WS, Roth MB: Distinct Functions of SR Proteins in Alternative Pre-mRNA Splicing. Science. 1993, 260: 219-222.

    CAS  PubMed  Google Scholar 

  54. Graveley BR: Sorting out the complexity of SR protein functions. RNA. 2000, 6 (9): 1197-1211. 10.1017/S1355838200000960.

    PubMed Central  CAS  PubMed  Google Scholar 

  55. Jacquenet S, Decimo D, Muriaux D, Darlix JL: Dual effect of the SR proteins ASF/SF2, SC35 and 9G8 on HIV-1 RNA splicing and virion production. Retrovirology. 2005, 2 (1): 33-10.1186/1742-4690-2-33.

    PubMed Central  PubMed  Google Scholar 

  56. Ropers D, Ayadi L, Gattoni R, Jacquenet S, Damier L, Branlant C, Stevenin J: Differential effects of the SR proteins 9G8, SC35, ASF/SF2 and SRp40 on the utilization of the A1 to A5 splicing sites of HIV-1 RNA. J Biol Chem. 2004, 279: 29963-29973. 10.1074/jbc.M404452200.

    CAS  PubMed  Google Scholar 

  57. Wentz MP, Moore BE, Cloyd MW, Berget SM, Donehower LA: A naturally arising mutation of a potential silencer of exon splicing in human immunodeficiency virus type 1 induces dominant aberrant splicing and arrests virus production. Journal of Virology. 1997, 71 (11): 8542-8551.

    PubMed Central  CAS  PubMed  Google Scholar 

  58. Liao HJ, Baker CC, Princler GL, Derse D: cis-Acting and trans-acting modulation of equine infectious anemia virus alternative RNA splicing. Virology. 2004, 323 (1): 131-140. 10.1016/j.virol.2003.12.028.

    CAS  PubMed  Google Scholar 

  59. Belshan M, Park GS, Bilodeau P, Stoltzfus CM, Carpenter S: Binding of equine infectious anemia virus rev to an exon splicing enhancer mediates alternative splicing and nuclear export of viral mRNAs. Mol Cell Biol. 2000, 20 (10): 3550-3557. 10.1128/MCB.20.10.3550-3557.2000.

    PubMed Central  CAS  PubMed  Google Scholar 

  60. Chung H, Derse D: Binding sites for Rev and ASF/SF2 map to a 55-nucleotide purine-rich exonic element in equine infectious anemia virus RNA. J Biol Chem. 2001, 276 (22): 18960-18967. 10.1074/jbc.M008996200.

    CAS  PubMed  Google Scholar 

  61. Gontarek RR, Derse D: Interactions among SR proteins, an exonic splicing enhancer, and a lentivirus Rev protein regulate alternative splicing. Mol Cell Biol. 1996, 16 (5): 2325-2331.

    PubMed Central  CAS  PubMed  Google Scholar 

  62. Cereseto A, Berneman Z, Koralnik I, Vaughn J, Franchini G, Klotman ME: Differential expression of alternatively spliced pX mRNAs in HTLV-I-infected cell lines. Leukemia. 1997, 11 (6): 866-870. 10.1038/sj.leu.2400665.

    CAS  PubMed  Google Scholar 

  63. Princler GL, Julias JG, Hughes SH, Derse D: Roles of viral and cellular proteins in the expression of alternatively spliced HTLV-1 pX mRNAs. Virology. 2003, 317 (1): 136-145. 10.1016/j.virol.2003.09.010.

    CAS  PubMed  Google Scholar 

  64. Kress E, Baydoun HH, Bex F, Gazzolo L, Duc Dodon M: Critical role of hnRNP A1 in HTLV-1 replication in human transformed T lymphocytes. Retrovirology. 2005, 2 (1): 8-10.1186/1742-4690-2-8.

    PubMed Central  PubMed  Google Scholar 

  65. Xu Y, Reddy TR, Fischer WH, Wong-Staal F: A Novel hnRNP Specifically Interacts with HIV-1 RRE RNA. Journal of Biomedical Science. 1996, 3 (2): 82-91. 10.1007/BF02255535.

    CAS  PubMed  Google Scholar 

  66. Oshima M, Odawara T, Hanaki K, Igarashi H, Yoshikura H: cis Elements required for high-level expression of unspliced Gag-containing message in Moloney murine leukemia virus. J Virol. 1998, 72 (8): 6414-6420.

    PubMed Central  CAS  PubMed  Google Scholar 

  67. Hoshi S, Odawara T, Oshima M, Kitamura Y, Takizawa H, Yoshikura H: cis-Elements involved in expression of unspliced RNA in Moloney murine leukemia virus. Biochem Biophys Res Commun. 2002, 290 (3): 1139-1144. 10.1006/bbrc.2001.6326.

    CAS  PubMed  Google Scholar 

  68. Hwang LS, Park J, Gilboa E: Role of intron-contained sequences in formation of moloney murine leukemia virus env mRNA. Mol Cell Biol. 1984, 4 (11): 2289-2297.

    PubMed Central  CAS  PubMed  Google Scholar 

  69. Aagaard L, Rasmussen SV, Mikkelsen JG, Pedersen FS: Efficient replication of full-length murine leukemia viruses modified at the dimer initiation site regions. Virology. 2004, 318 (1): 360-370. 10.1016/j.virol.2003.09.008.

    CAS  PubMed  Google Scholar 

  70. Lee JT, Yu SS, Han E, Kim S, Kim S: Engineering the splice acceptor for improved gene expression and viral titer in an MLV-based retroviral vector. Gene Ther. 2004, 11 (1): 94-99. 10.1038/sj.gt.3302138.

    CAS  PubMed  Google Scholar 

  71. Zheng YH, Yu HF, Peterlin BM: Human p32 protein relieves a post-transcriptional block to HIV replication in murine cells. Nat Cell Biol. 2003, 5 (7): 611-618. 10.1038/ncb1000.

    CAS  PubMed  Google Scholar 

  72. Das AT, Klaver B, Berkhout B: A hairpin structure in the R region of the human immunodeficiency virus type 1 RNA genome is instrumental in polyadenylation site selection. Journal of Virology. 1999, 73 (1): 81-91.

    PubMed Central  CAS  PubMed  Google Scholar 

  73. Klasens BI, Thiesen M, Virtanen A, Berkhout B: The ability of the HIV-1 AAUAAA signal to bind polyadenylation factors is controlled by local RNA structure. Nucleic Acids Research. 1999, 27 (2): 446-454. 10.1093/nar/27.2.446.

    PubMed Central  CAS  PubMed  Google Scholar 

  74. Vagner S, Ruegsegger U, Gunderson SI, Keller W, Mattaj IW: Position-dependent inhibition of the cleavage step of pre-mRNA 3'-end processing by U1 snRNP. Rna-A Publication of the Rna Society. 2000, 6 (2): 178-188.

    CAS  Google Scholar 

  75. Ashe MP, Furger A, Proudfoot NJ: Stem-loop 1 of the U1 snRNP plays a critical role in the suppression of HIV-1 polyadenylation. Rna-A Publication of the Rna Society. 2000, 6 (2): 170-177.

    CAS  Google Scholar 

  76. Ashe MP, Pearson LH, Proudfoot NJ: The HIV-1 5' LTR poly(A) site is inactivated by U1 snRNP interaction with the downstream major splice donor site. EMBO Journal. 1997, 16 (18): 5752-5763. 10.1093/emboj/16.18.5752.

    PubMed Central  CAS  PubMed  Google Scholar 

  77. Gunderson SI, Polycarpou-Schwarz M, Mattaj IW: U1 snRNP inhibits pre-mRNA polyadenylation through a direct interaction between U1 70K and poly(A) polymerase. Molecular Cell. 1998, 1 (2): 255-264. 10.1016/S1097-2765(00)80026-X.

    CAS  PubMed  Google Scholar 

  78. Pongoski J, Asai K, Cochrane A: Positive and Negative Modulation of Human Immunodeficiency Virus Type 1 Rev Function by cis and trans Regulators of Viral RNA Splicing. J Virol. 2002, 76 (10): 5108-5120. 10.1128/JVI.76.10.5108-5120.2002.

    PubMed Central  CAS  PubMed  Google Scholar 

  79. McLaren M, Asai K, Cochrane A: A novel function for Sam68: Enhancement of HIV-1 RNA 3' end processing. RNA. 2004, 10: 1119-1129. 10.1261/rna.5263904.

    PubMed Central  CAS  PubMed  Google Scholar 

  80. Reddy TR, Xu W, Wong-Staal F: General effect of Sam68 on Rev/Rex regulated expression of complex retroviruses. Oncogene. 2000, 19: 4071-4074. 10.1038/sj.onc.1203749.

    CAS  PubMed  Google Scholar 

  81. Reddy T, Xu W, Mau J, Goodwin C, Suhasini M, Tang H, Frimpong K, Rose D, Wong-Staal F: Inhibition of HIV replication by dominant negative mutants of Sam68, a functional homolog of HIV-1 Rev. Nature Medicine. 1999, 5: 635-642. 10.1038/9479.

    CAS  PubMed  Google Scholar 

  82. Reddy TR, Suhasini M, Xu W, Yeh LY, Yang JP, Wu J, Artzt K, Wong-Staal F: A role for KH domain proteins (Sam68-like mammalian proteins and quaking proteins) in the post-transcriptional regulation of HIV replication. Journal of Biological Chemistry. 2002, 277 (8): 5778-5784. 10.1074/jbc.M106836200.

    CAS  PubMed  Google Scholar 

  83. Soros V, Valderrarama Carvajal H, Richard S, Cochrane A: Inhibition of Human Immunodeficiency Virus Type 1 Rev Function by a Dominant-Negative Mutant of Sam68 through Sequestration of Unspliced RNA at Perinuclear Bundles. J Virol. 2001, 75: 8203-8215. 10.1128/JVI.75.17.8203-8215.2001.

    PubMed Central  CAS  PubMed  Google Scholar 

  84. Modem S, Badri KR, Holland TC, Reddy TR: Sam68 is absolutely required for Rev function and HIV-1 production. Nucleic Acids Research. 2005, 33 (3): 873-879. 10.1093/nar/gki231.

    PubMed Central  CAS  PubMed  Google Scholar 

  85. Mouland AJ, Coady M, Yao XJ, Cohen EA: Hypophosphorylation of poly(A) polymerase and increased polyadenylation activity are associated with human immunodeficiency virus type 1 Vpr expression. Virology. 2002, 292 (2): 321-330. 10.1006/viro.2001.1261.

    CAS  PubMed  Google Scholar 

  86. Calzado MA, Sancho R, Munoz E: Human immunodeficiency virus type 1 Tat increases the expression of cleavage and polyadenylation specificity factor 73-kilodalton subunit modulating cellular and viral expression. Journal of Virology. 2004, 78 (13): 6846-6854. 10.1128/JVI.78.13.6846-6854.2004.

    PubMed Central  CAS  PubMed  Google Scholar 

  87. Furger A, Monks J, Proudfoot NJ: The retroviruses human immunodeficiency virus type 1 and Moloney murine leukemia virus adopt radically different strategies to regulate promoter-proximal polyadenylation. J Virol. 2001, 75 (23): 11735-11746. 10.1128/JVI.75.23.11735-11746.2001.

    PubMed Central  CAS  PubMed  Google Scholar 

  88. Herman SA, Coffin JM: Differential transcription from the long terminal repeats of integrated avian leukosis virus DNA. J Virol. 1986, 60 (2): 497-505.

    PubMed Central  CAS  PubMed  Google Scholar 

  89. Miller JT, Stoltzfus CM: Two distant upstream regions containing cis-acting signals regulating splicing facilitate 3'-end processing of avian sarcoma virus RNA. Journal of virology. 1992, 66 (7): 4242-4251.

    PubMed Central  CAS  PubMed  Google Scholar 

  90. Fogel BL, McNally LM, McNally MT: Efficient polyadenylation of Rous sarcoma virus RNA requires the negative regulator of splicing element. Nucleic Acids Research. 2002, 30 (3): 810-817. 10.1093/nar/30.3.810.

    PubMed Central  CAS  PubMed  Google Scholar 

  91. Polony TS, Bowers SJ, Neiman PE, Beemon KL: Silent point mutation in an avian retrovirus RNA processing element promotes c-myb-associated short-latency lymphomas. J Virol. 2003, 77 (17): 9378-9387. 10.1128/JVI.77.17.9378-9387.2003.

    PubMed Central  CAS  PubMed  Google Scholar 

  92. Dimaano C, Ullman KS: Nucleocytoplasmic transport: integrating mRNA production and turnover with export through the nuclear pore. Mol Cell Biol. 2004, 24 (8): 3069-3076. 10.1128/MCB.24.8.3069-3076.2004.

    PubMed Central  CAS  PubMed  Google Scholar 

  93. Cullen BR: Nuclear RNA export. Journal of Cell Science. 2003, 116 (Pt 4): 587-597. 10.1242/jcs.00268.

    PubMed  Google Scholar 

  94. Cullen B: HIV-1 Auxiliary Proteins: Making Connections in a Dying Cell. Cell. 1998, 93: 685-692. 10.1016/S0092-8674(00)81431-2.

    CAS  PubMed  Google Scholar 

  95. Hope TJ: The ins and outs of HIV Rev. Archives of Biochemistry & Biophysics. 1999, 365: 186-191. 10.1006/abbi.1999.1207.

    CAS  Google Scholar 

  96. Pollard V, Malim M: The HIV-1 Rev Protein. Annual Review of Microbiology. 1998, 52: 491-532. 10.1146/annurev.micro.52.1.491.

    CAS  PubMed  Google Scholar 

  97. Chang DD, Sharp PA: Regulation by HIV Rev depends upon recognition of splice sites. Cell. 1989, 59: 789-795. 10.1016/0092-8674(89)90602-8.

    CAS  PubMed  Google Scholar 

  98. Brighty DW, Rosenberg M: A cis-acting repressive sequence that overlaps the Rev-responsive element of human immunodeficiency virus type 1 regulates nuclear retention of env mRNAs independently of known splicing signals. Proceedings of the National Academy of Sciences of the United States if America. 1994, 91: 8314-8318.

    CAS  Google Scholar 

  99. Cochrane AW, Jones KS, Beidas S, Dillon PJ, Skalka AM, Rosen CA: Identification and characterization of intragenic sequences which repress human immunodeficiency virus structural gene expression. Journal of Virology. 1991, 65: 5305-5313.

    PubMed Central  CAS  PubMed  Google Scholar 

  100. Keller R, Montagnier L, Cordonnier A: Characterization of a Nuclear Retention Sequence within the 3' Region of the HIV-2 Envelope Gene. Virology. 1993, 192: 33-37. 10.1006/viro.1993.1005.

    CAS  PubMed  Google Scholar 

  101. Maldarelli F, Martin MA, Strebel K: Identification of post-transcriptionally active inhibitory sequences in human immunodeficiency virus type I RNA: novel level of gene regulation. Journal of Virology. 1991, 65: 5732-5743.

    PubMed Central  CAS  PubMed  Google Scholar 

  102. Mikaelian I, Krieg M, Gait M, Karn J: Interactions of INS (CRS) Elements and the Splicing Machinery Regulate the Production of Rev-responsive mRNAs. Journal of Molecular Biology. 1996, 257: 246-264. 10.1006/jmbi.1996.0160.

    CAS  PubMed  Google Scholar 

  103. Schwartz S, Felber BK, Pavlakis GN: Distinct RNA sequences in the gag region of human immunodeficiency virus type 1 decreases RNA stability and inhibit expression in the absence of Rev protein. Journal of Virology. 1992, 66: 150-159.

    PubMed Central  CAS  PubMed  Google Scholar 

  104. Schwartz S, Campbell M, Nasioulas G, Harrison J, Felber B, Pavlakis G: Mutational inactivation of an inhibitory sequence in human immunodeficiency virus type 1 results in Rev-independent gag expression. Journal of Virology. 1992, 66: 7176-7182.

    PubMed Central  CAS  PubMed  Google Scholar 

  105. Seguin B, Staffa A, Cochrane A: Control of HIV-1 RNA Metabolism: The Role of Splice Sites and Intron Sequences in Unspliced Viral RNA Subcellular Distribution. Journal of Virology. 1998, 72: 9503-9513.

    PubMed Central  CAS  PubMed  Google Scholar 

  106. Nasioulas G, Zolotukhin A, Tabernero C, Solomin L, Cunningham C, Pavlakis G, Felber B: Elements Distinct from Human Immunodeficiency Virus Type 1 Splice Sites Are Responsible for the Rev Dependence of env mRNA. Journal of Virology. 1994, 68: 2986-2993.

    PubMed Central  CAS  PubMed  Google Scholar 

  107. Graf M, Bojak A, Deml L, Bieler H, Wolf H, Wagner R: Concerted Action of Multiple cis-acting Sequences Is Required for Rev Dependence of Late Human Immunodeficiency Virus Type 1 Gene Expression. J Virol. 2000, 74: 10822-10826. 10.1128/JVI.74.22.10822-10826.2000.

    PubMed Central  CAS  PubMed  Google Scholar 

  108. Iacampo S, Cochrane A: Human Immunodeficiency Virus Type 1 Rev Function Requires Continued Synthesis of Its Target mRNA. Journal of Virology. 1996, 70: 8332-8339.

    PubMed Central  CAS  PubMed  Google Scholar 

  109. Huang Y, Carmichael GC: Role of polyadenylation in nucleocytoplasmic transport of mRNA. Molecular & Cellular Biology. 1996, 16 (4): 1534-1542.

    CAS  Google Scholar 

  110. Maniatis T, Reed R: An extensive network of coupling among gene expression machines. Nature. 2002, 416 (6880): 499-506. 10.1038/416499a.

    CAS  PubMed  Google Scholar 

  111. Reed R: Coupling transcription, splicing and mRNA export. Current Opinion in Cell Biology. 2003, 15 (3): 326-331. 10.1016/S0955-0674(03)00048-6.

    CAS  PubMed  Google Scholar 

  112. Vasudevan S, Peltz SW: Nuclear mRNA surveillance. Current Opinion in Cell Biology. 2003, 15 (3): 332-337. 10.1016/S0955-0674(03)00051-6.

    CAS  PubMed  Google Scholar 

  113. Yedavalli VS, Neuveut C, Chi YH, Kleiman L, Jeang KT: Requirement of DDX3 DEAD box RNA helicase for HIV-1 Rev-RRE export function. Cell. 2004, 119 (3): 381-392. 10.1016/j.cell.2004.09.029.

    CAS  PubMed  Google Scholar 

  114. Fang J, Kubota S, Yang B, Zhou N, Zhang H, Godbout R, Pomerantz RJ: A DEAD box protein facilitates HIV-1 replication as a cellular co-factor of Rev. Virology. 2004, 330 (2): 471-480. 10.1016/j.virol.2004.09.039.

    CAS  PubMed  Google Scholar 

  115. Yu Z, Sanchez-Velar N, Catrina IE, Kittler EL, Udofia EB, Zapp ML: The cellular HIV-1 Rev cofactor hRIP is required for viral replication. Proceedings of the National Academy of Sciences of the United States of America. 2005, 102 (11): 4027-4032. 10.1073/pnas.0408889102.

    PubMed Central  CAS  PubMed  Google Scholar 

  116. Sanchez-Velas N, Udofia I, Yu Z, Zapp M: hRIP, a cellular cofactor for Rev function, promotes release of HIV-1 RNAs from the perinuclear region. Genes & Dev. 2004, 18: 23-34. 10.1101/gad.1149704.

    Google Scholar 

  117. Bray M, Prasad S, Dubay JW, Hunter E, Jeang KT, Rekosh D, Hammarskjold ML: A small element from the Mason-Pfizer monkey virus genome makes human immunodeficiency virus type 1 expression and replication Rev- independent. Proc Natl Acad Sci U S A. 1994, 91 (4): 1256-1260.

    PubMed Central  CAS  PubMed  Google Scholar 

  118. Ernst RK, Bray M, Rekosh D, Hammarskjold ML: A structured retroviral RNA element that mediates nucleocytoplasmic export of intron-containing RNA. Molecular & Cellular Biology. 1997, 17 (1): 135-144.

    CAS  Google Scholar 

  119. Hammarskjold ML, Li H, Rekosh D, Prasad S: Human Immunodeficiency Virus env Expression Becomes Rev-Independent if the env Region Is Not Defined as an Intron. Journal of Virology. 1994, 68: 951-958.

    PubMed Central  CAS  PubMed  Google Scholar 

  120. Zolotukhin AS, Valentin A, Pavlakis GN, Felber BK: Continuous propagation of RRE(-) and Rev(-)RRE(-) human immunodeficiency virus type 1 molecular clones containing a cis-acting element of simian retrovirus type 1 in human peripheral blood lymphocytes. J Virol. 1994, 68 (12): 7944-7952.

    PubMed Central  CAS  PubMed  Google Scholar 

  121. Pasquinelli AE, Ernst RK, Lund E, Grimm C, Zapp ML, Rekosh D, Hammarskjold ML, Dahlberg JE: The constitutive transport element (CTE) of Mason-Pfizer monkey virus (MPMV) accesses a cellular mRNA export pathway. Embo J. 1997, 16 (24): 7500-7510. 10.1093/emboj/16.24.7500.

    PubMed Central  CAS  PubMed  Google Scholar 

  122. Saavedra C, Felber B, Izaurralde E: The simian retrovirus-1 constitutive transport element, unlike the HIV-1 RRE, uses factors required for cellular mRNA export. Curr Biol. 1997, 7 (9): 619-628. 10.1016/S0960-9822(06)00288-0.

    CAS  PubMed  Google Scholar 

  123. Gruter P, Tabernero C, von Kobbe C, Schmitt C, Saavedra C, Bachi A, Wilm M, Felber B, Izauralde E: TAP, the Human Homolog of Mex67p, Mediates CTE-Dependent RNA Export from the Nucleus. Molecular Cell. 1998, 1: 649-659. 10.1016/S1097-2765(00)80065-9.

    CAS  PubMed  Google Scholar 

  124. Ogert RA, Lee LH, Beemon KL: Avian retroviral RNA element promotes unspliced RNA accumulation in the cytoplasm. Journal of Virology. 1996, 70 (6): 3834-3843.

    PubMed Central  CAS  PubMed  Google Scholar 

  125. Yang J, Cullen BR: Structural and functional analysis of the avian leukemia virus constitutive transport element. Rna. 1999, 5 (12): 1645-1655. 10.1017/S1355838299991616.

    PubMed Central  CAS  PubMed  Google Scholar 

  126. Simpson SB, Zhang L, Craven RC, Stoltzfus CM: Rous sarcoma virus direct repeat cis elements exert effects at several points in the virus life cycle. J Virol. 1997, 71 (12): 9150-9156.

    PubMed Central  CAS  PubMed  Google Scholar 

  127. Simpson SB, Guo W, Winistorfer SC, Craven RC, Stoltzfus CM: The upstream, direct repeat sequence of Prague A Rous sarcoma virus is deficient in mediating efficient Gag assembly and particle release. Virology. 1998, 247 (1): 86-96. 10.1006/viro.1998.9233.

    CAS  PubMed  Google Scholar 

  128. Paca RE, Ogert RA, Hibbert CS, Izaurralde E, Beemon KL: Rous sarcoma virus DR posttranscriptional elements use a novel RNA export pathway. Journal of Virology. 2000, 74 (20): 9507-9514. 10.1128/JVI.74.20.9507-9514.2000.

    PubMed Central  CAS  PubMed  Google Scholar 

  129. Tange TO, Nott A, Moore MJ: The ever-increasing complexities of the exon junction complex. Curr Opin Cell Biol. 2004, 16 (3): 279-284. 10.1016/j.ceb.2004.03.012.

    CAS  PubMed  Google Scholar 

  130. Berberich SL, Macias M, Zhang L, Turek LP, Stoltzfus CM: Comparison of Rous sarcoma virus RNA processing in chicken and mouse fibroblasts: evidence for double-spliced RNA in nonpermissive mouse cells [published erratum appears in J Virol 1990 Dec;64(12):6360]. Journal of Virology. 1990, 64 (9): 4313-4320.

    PubMed Central  CAS  PubMed  Google Scholar 

  131. Vogt VM, Bruckenstein DA, Bell AP: Avian sarcoma virus gag precursor polypeptide is not processed in mammalian cells. J Virol. 1982, 44 (2): 725-730.

    PubMed Central  CAS  PubMed  Google Scholar 

  132. Swanson CM, Puffer BA, Ahmad KM, Doms RW, Malim MH: Retroviral mRNA nuclear export elements regulate protein function and virion assembly. Embo J. 2004, 23 (13): 2632-2640. 10.1038/sj.emboj.7600270.

    PubMed Central  CAS  PubMed  Google Scholar 

  133. Beriault V, Clement JF, Levesque K, Lebel C, Yong X, Chabot B, Cohen EA, Cochrane AW, Rigby WF, Mouland AJ: A late role for the association of hnRNP A2 with the HIV-1 hnRNP A2 response elements in genomic RNA, Gag, and Vpr localization. J Biol Chem. 2004, 279 (42): 44141-44153. 10.1074/jbc.M404691200.

    CAS  PubMed  Google Scholar 

  134. Butsch M, Boris-Lawrie K: Destiny of unspliced retroviral RNA: ribosome and/or virion?. J Virol. 2002, 76 (7): 3089-3094. 10.1128/JVI.76.7.3089-3094.2002.

    PubMed Central  CAS  PubMed  Google Scholar 

  135. Kaye JF, Lever AM: Human immunodeficiency virus types 1 and 2 differ in the predominant mechanism used for selection of genomic RNA for encapsidation. Journal of Virology. 1999, 73 (4): 3023-3031.

    PubMed Central  CAS  PubMed  Google Scholar 

  136. Poon DT, Chertova EN, Ott DE: Human Immunodeficiency Virus Type 1 Preferentially Encapsidates Genomic RNAs That Encode Pr55(Gag): Functional Linkage between Translation and RNA Packaging. Virology. 2002, 293 (2): 368-378. 10.1006/viro.2001.1283.

    CAS  PubMed  Google Scholar 

  137. Levin JG, Rosenak MJ: Synthesis of murine leukemia virus proteins associated with virions assembled in actinomycin D-treated cells: evidence for persistence of viral messenger RNA. Proc Natl Acad Sci U S A. 1976, 73 (4): 1154-1158.

    PubMed Central  CAS  PubMed  Google Scholar 

  138. Butsch M, Boris-Lawrie K: Translation is not required To generate virion precursor RNA in human immunodeficiency virus type 1-infected T cells. J Virol. 2000, 74 (24): 11531-11537. 10.1128/JVI.74.24.11531-11537.2000.

    PubMed Central  CAS  PubMed  Google Scholar 

  139. Poole E, Strappe P, Mok HP, Hicks R, Lever AM: HIV-1 Gag-RNA interaction occurs at a perinuclear/centrosomal site; analysis by confocal microscopy and FRET. Traffic. 2005, 6 (9): 741-755. 10.1111/j.1600-0854.2005.00312.x.

    CAS  PubMed  Google Scholar 

  140. Muriaux D, Mirro J, Harvin D, Rein A: RNA is a structural element in retrovirus particles. Proc Natl Acad Sci U S A. 2001, 98 (9): 5246-5251. 10.1073/pnas.091000398.

    PubMed Central  CAS  PubMed  Google Scholar 

  141. Ott DE, Coren LV, Gagliardi TD: Redundant roles for nucleocapsid and matrix RNA-binding sequences in human immunodeficiency virus type 1 assembly. J Virol. 2005, 79 (22): 13839-13847. 10.1128/JVI.79.22.13839-13847.2005.

    PubMed Central  CAS  PubMed  Google Scholar 

  142. Sfakianos JN, LaCasse RA, Hunter E: The M-PMV cytoplasmic targeting-retention signal directs nascent Gag polypeptides to a pericentriolar region of the cell. Traffic. 2003, 4 (10): 660-670. 10.1034/j.1600-0854.2003.00125.x.

    CAS  PubMed  Google Scholar 

  143. Sfakianos JN, Hunter E: M-PMV capsid transport is mediated by Env/Gag interactions at the pericentriolar recycling endosome. Traffic. 2003, 4 (10): 671-680. 10.1034/j.1600-0854.2003.00126.x.

    CAS  PubMed  Google Scholar 

  144. Derdowski A, Ding L, Spearman P: A novel fluorescence resonance energy transfer assay demonstrates that the human immunodeficiency virus type 1 Pr55Gag I domain mediates Gag-Gag interactions. J Virol. 2004, 78 (3): 1230-1242. 10.1128/JVI.78.3.1230-1242.2004.

    PubMed Central  CAS  PubMed  Google Scholar 

  145. Larson DR, Ma YM, Vogt VM, Webb WW: Direct measurement of Gag-Gag interaction during retrovirus assembly with FRET and fluorescence correlation spectroscopy. J Cell Biol. 2003, 162 (7): 1233-1244. 10.1083/jcb.200303200.

    PubMed Central  CAS  PubMed  Google Scholar 

  146. Mouland AJ, Cohen A, DesGroseillers L: Trafficking of HIV-1 RNA: Recent Progress Involving Host Cell RNA-Binding Proteins. Current Genomics. 2003, 4 (3): 196-10.2174/1389202033490402.

    Google Scholar 

  147. Fusco D, Accornero N, Lavoie B, Shenoy SM, Blanchard JM, Singer RH, Bertrand E: Single mRNA molecules demonstrate probabilistic movement in living Mammalian cells. Current Biology. 2003, 13 (2): 161-167. 10.1016/S0960-9822(02)01436-7.

    CAS  PubMed  Google Scholar 

  148. Basyuk E, Galli T, Mougel M, Blanchard JM, Sitbon M, Bertrand E: Retroviral genomic RNAs are transported to the plasma membrane by endosomal vesicles. Developmental Cell. 2003, 5 (1): 161-174. 10.1016/S1534-5807(03)00188-6.

    CAS  PubMed  Google Scholar 

  149. Basyuk E, Boulon S, Skou Pedersen F, Bertrand E, Vestergaard Rasmussen S: The packaging signal of MLV is an integrated module that mediates intracellular transport of genomic RNAs. J Mol Biol. 2005, 354 (2): 330-339. 10.1016/j.jmb.2005.09.071.

    CAS  PubMed  Google Scholar 

  150. Hoek KS, Kidd GJ, Carson JH, Smith R: hnRNP A2 selectively binds the cytoplasmic transport sequence of myelin basic protein mRNA. Biochemistry. 1998, 37 (19): 7021-7029. 10.1021/bi9800247.

    CAS  PubMed  Google Scholar 

  151. Mouland AJ, Xu H, Cui H, Krueger W, Munro TP, Prasol M, Mercier J, Rekosh D, Smith R, Barbarese E, Cohen EA, Carson JH: RNA trafficking signals in human immunodeficiency virus type 1. Mol Cell Biol. 2001, 21 (6): 2133-2143. 10.1128/MCB.21.6.2133-2143.2001.

    PubMed Central  CAS  PubMed  Google Scholar 

  152. Kiebler MA, DesGroseillers L: Molecular insights into mRNA transport and local translation in the mammalian nervous system. Neuron. 2000, 25 (1): 19-28. 10.1016/S0896-6273(00)80868-5.

    CAS  PubMed  Google Scholar 

  153. Barbarese E, Koppel DE, Deutscher MP, Smith CL, Ainger K, Morgan F, Carson JH: Protein translation components are colocalized in granules in oligodendrocytes. J Cell Sci. 1995, 108 (Pt 8): 2781-2790.

    CAS  PubMed  Google Scholar 

  154. Shav-Tal Y, Singer RH: RNA localization. J Cell Sci. 2005, 118 (Pt 18): 4077-4081. 10.1242/jcs.02543.

    CAS  PubMed  Google Scholar 

  155. Mingle LA, Okuhama NN, Shi J, Singer RH, Condeelis J, Liu G: Localization of all seven messenger RNAs for the actin-polymerization nucleator Arp2/3 complex in the protrusions of fibroblasts. J Cell Sci. 2005, 118 (Pt 11): 2425-2433. 10.1242/jcs.02371.

    PubMed Central  CAS  PubMed  Google Scholar 

  156. Reddy TR, Tang H, Xu W, Wong-Staal F: Sam68, RNA helicase A and Tap cooperate in the post-transcriptional regulation of human immunodeficiency virus and type D retroviral mRNA. Oncogene. 2000, 19 (32): 3570-3575. 10.1038/sj.onc.1203676.

    CAS  PubMed  Google Scholar 

  157. Li J, Tang H, Mullen TM, Westberg C, Reddy TR, Rose DW, Wong-Staal F: A role for RNA helicase A in post-transcriptional regulation of HIV type 1. Proc Natl Acad Sci U S A. 1999, 96 (2): 709-714. 10.1073/pnas.96.2.709.

    PubMed Central  CAS  PubMed  Google Scholar 

  158. Tang H, Wong-Staal F: Specific interaction between RNA helicase A and Tap, two cellular proteins that bind to the constitutive transport element of type D retrovirus. J Biol Chem. 2000, 275 (42): 32694-32700. 10.1074/jbc.M003933200.

    CAS  PubMed  Google Scholar 

  159. Krishnan V, Zeichner SL: Alterations in the expression of DEAD-box and other RNA binding proteins during HIV-1 replication. Retrovirology. 2004, 1 (1): 42-10.1186/1742-4690-1-42.

    PubMed Central  PubMed  Google Scholar 

  160. Kanai Y, Dohmae N, Hirokawa N: Kinesin transports RNA: isolation and characterization of an RNA-transporting granule. Neuron. 2004, 43 (4): 513-525. 10.1016/j.neuron.2004.07.022.

    CAS  PubMed  Google Scholar 

  161. Reichert VL, Le Hir H, Jurica MS, Moore MJ: 5' exon interactions within the human spliceosome establish a framework for exon junction complex structure and assembly. Genes & Development. 2002, 16 (21): 2778-2791. 10.1101/gad.1030602.

    CAS  Google Scholar 

  162. Monshausen M, Gehring NH, Kosik KS: The mammalian RNA-binding protein Staufen2 links nuclear and cytoplasmic RNA processing pathways in neurons. Neuromolecular Med. 2004, 6 (2-3): 127-144. 10.1385/NMM:6:2-3:127.

    CAS  PubMed  Google Scholar 

  163. Shyu AB, Wilkinson MF: The double lives of shuttling mRNA binding proteins. Cell. 2000, 102 (2): 135-138. 10.1016/S0092-8674(00)00018-0.

    CAS  PubMed  Google Scholar 

  164. Stoltzfus CM, Madsen JM: Role of Viral Splicing Elements and Cellular RNA Binding Proteins in Regulation of HIV-1 Alternative RNA Splicing. Curr HIV Res. 2006, 4 (1): 43-55. 10.2174/157016206775197655.

    CAS  PubMed  Google Scholar 

  165. Najera I, Krieg M, Karn J: Synergistic stimulation of HIV-1 rev-dependent export of unspliced mRNA to the cytoplasm by hnRNP A1. J Mol Biol. 1999, 285 (5): 1951-1964. 10.1006/jmbi.1998.2473.

    CAS  PubMed  Google Scholar 

  166. Ainger K, Avossa D, Diana AS, Barry C, Barbarese E, Carson JH: Transport and localization elements in myelin basic protein mRNA. J Cell Biol. 1997, 138 (5): 1077-1087. 10.1083/jcb.138.5.1077.

    PubMed Central  CAS  PubMed  Google Scholar 

  167. Kosturko LD, Maggipinto MJ, D'Sa C, Carson JH, Barbarese E: The Microtubule-associated Protein, TOG, Binds to the RNA Trafficking Protein, hnRNP A2. Mol Biol Cell. 2005

    Google Scholar 

  168. Tang Y, Winkler U, Freed EO, Torrey TA, Kim W, Li H, Goff SP, Morse HC: Cellular motor protein KIF-4 associates with retroviral Gag. J Virol. 1999, 73 (12): 10508-10513.

    PubMed Central  CAS  PubMed  Google Scholar 

  169. Kim W, Tang Y, Okada Y, Torrey TA, Chattopadhyay SK, Pfleiderer M, Falkner FG, Dorner F, Choi W, Hirokawa N, Morse HC: Binding of murine leukemia virus Gag polyproteins to KIF4, a microtubule-based motor protein. J Virol. 1998, 72 (8): 6898-6901.

    PubMed Central  CAS  PubMed  Google Scholar 

  170. Kohrmann M, Luo M, Kaether C, DesGroseillers L, Dotti CG, Kiebler MA: Microtubule-dependent recruitment of Staufen-green fluorescent protein into large RNA-containing granules and subsequent dendritic transport in living hippocampal neurons. Mol Biol Cell. 1999, 10 (9): 2945-2953.

    PubMed Central  CAS  PubMed  Google Scholar 

  171. Mallardo M, Deitinghoff A, Muller J, Goetze B, Macchi P, Peters C, Kiebler MA: From the Cover: Isolation and characterization of Staufen-containing ribonucleoprotein particles from rat brain. Proc Natl Acad Sci U S A. 2003, 100 (4): 2100-2105. 10.1073/pnas.0334355100.

    PubMed Central  CAS  PubMed  Google Scholar 

  172. Chatel-Chaix L, Clément JF, Martel C, Bériault V, Gatignol A, DesGroseillers L, Mouland AJ: Staufen is part of the HIV-1 Gag Ribonucleoprotein complex and is involved in the generation of infectious viral particles. Molecular and Cellular Biology. 2004, 24 (7): 2637-2648. 10.1128/MCB.24.7.2637-2648.2004.

    PubMed Central  CAS  PubMed  Google Scholar 

  173. Ferrandon D, Koch I, Westhof E, Nusslein-Volhard C: RNA-RNA interaction is required for the formation of specific bicoid mRNA 3' UTR-STAUFEN ribonucleoprotein particles. EMBO Journal. 1997, 16 (7): 1751-1758. 10.1093/emboj/16.7.1751.

    PubMed Central  CAS  PubMed  Google Scholar 

  174. Paillart JC, Shehu-Xhilaga M, Marquet R, Mak J: Dimerization of retroviral RNA genomes: an inseparable pair. Nat Rev Microbiol. 2004, 2 (6): 461-472. 10.1038/nrmicro903.

    CAS  PubMed  Google Scholar 

  175. Mouland AJ, Mercier J, Luo M, Bernier L, DesGroseillers L, Cohen EA: The double-stranded RNA-binding protein Staufen is incorporated in human immunodeficiency virus type 1: evidence for a role in genomic RNA encapsidation. J Virol. 2000, 74 (12): 5441-5451. 10.1128/JVI.74.12.5441-5451.2000.

    PubMed Central  CAS  PubMed  Google Scholar 

  176. Soret J, Bakkour N, Maire S, Durand S, Zekri L, Gabut M, Fic W, Divita G, Rivalle C, Dauzonne D, Nguyen CH, Jeanteur P, Tazi J: Selective modification of alternative splicing by indole derivatives that target serine-arginine-rich protein splicing factors. Proc Natl Acad Sci U S A. 2005, 102 (24): 8764-8769. 10.1073/pnas.0409829102.

    PubMed Central  CAS  PubMed  Google Scholar 

  177. Daniel R, Pomerantz RJ: ATM: HIV-1's Achilles heel?. Nat Cell Biol. 2005, 7 (5): 452-453. 10.1038/ncb0505-452.

    CAS  PubMed  Google Scholar 

  178. Hofmann W, Reichart B, Ewald A, Muller E, Schmitt I, Stauber RH, Lottspeich F, Jockusch BM, Scheer U, Hauber J, Dabauvalle MC: Cofactor requirements for nuclear export of Rev response element (RRE)- and constitutive transport element (CTE)-containing retroviral RNAs. An unexpected role for actin. J Cell Biol. 2001, 152 (5): 895-910. 10.1083/jcb.152.5.895.

    PubMed Central  CAS  PubMed  Google Scholar 

  179. Liu B, Dai R, Tian CJ, Dawson L, Gorelick R, Yu XF: Interaction of the human immunodeficiency virus type 1 nucleocapsid with actin. J Virol. 1999, 73 (4): 2901-2908.

    PubMed Central  CAS  PubMed  Google Scholar 

  180. Wilk T, Gowen B, Fuller SD: Actin associates with the nucleocapsid domain of the human immunodeficiency virus Gag polyprotein. J Virol. 1999, 73 (3): 1931-1940.

    PubMed Central  CAS  PubMed  Google Scholar 

  181. Ott DE, Coren LV, Johnson DG, Kane BP, Sowder RC, Kim YD, Fisher RJ, Zhou XZ, Lu KP, Henderson LE: Actin-binding cellular proteins inside human immunodeficiency virus type 1. Virology. 2000, 266 (1): 42-51. 10.1006/viro.1999.0075.

    CAS  PubMed  Google Scholar 

  182. Cimarelli A, Luban J: Translation elongation factor 1-alpha interacts specifically with the human immunodeficiency virus type 1 Gag polyprotein. J Virol. 1999, 73 (7): 5388-5401.

    PubMed Central  CAS  PubMed  Google Scholar 

  183. Bevec D, Jaksche H, Oft M, Wohl T, Himmelspach M, Pacher A, Schebesta M, Koettnitz K, Dobrovnik M, Csonga R, Lottspeich F, Hauber J: Inhibition of HIV-1 replication in lymphocytes by mutants of the Rev cofactor eIF-5A. Science. 1996, 271 (5257): 1858-1860.

    CAS  PubMed  Google Scholar 

  184. Gurer C, Cimarelli A, Luban J: Specific incorporation of heat shock protein 70 family members into primate lentiviral virions. J Virol. 2002, 76 (9): 4666-4670. 10.1128/JVI.76.9.4666-4670.2002.

    PubMed Central  CAS  PubMed  Google Scholar 

  185. Gurer C, Hoglund A, Hoglund S, Luban J: ATPgammaS disrupts human immunodeficiency virus type 1 virion core integrity. J Virol. 2005, 79 (9): 5557-5567. 10.1128/JVI.79.9.5557-5567.2005.

    PubMed Central  CAS  PubMed  Google Scholar 

  186. Agostini I, Popov S, Li J, Dubrovsky L, Hao T, Bukrinsky M: Heat-shock protein 70 can replace viral protein R of HIV-1 during nuclear import of the viral preintegration complex. Exp Cell Res. 2000, 259 (2): 398-403. 10.1006/excr.2000.4992.

    CAS  PubMed  Google Scholar 

  187. Iordanskiy S, Zhao Y, Dubrovsky L, Iordanskaya T, Chen M, Liang D, Bukrinsky M: Heat shock protein 70 protects cells from cell cycle arrest and apoptosis induced by human immunodeficiency virus type 1 viral protein R. J Virol. 2004, 78 (18): 9697-9704. 10.1128/JVI.78.18.9697-9704.2004.

    PubMed Central  CAS  PubMed  Google Scholar 

  188. Bacharach E, Gonsky J, Alin K, Orlova M, Goff SP: The carboxy-terminal fragment of nucleolin interacts with the nucleocapsid domain of retroviral gag proteins and inhibits virion assembly. J Virol. 2000, 74 (23): 11027-11039. 10.1128/JVI.74.23.11027-11039.2000.

    PubMed Central  CAS  PubMed  Google Scholar 

  189. Ueno T, Tokunaga K, Sawa H, Maeda M, Chiba J, Kojima A, Hasegawa H, Shoya Y, Sata T, Kurata T, Takahashi H: Nucleolin and the packaging signal, psi, promote the budding of human immunodeficiency virus type-1 (HIV-1). Microbiol Immunol. 2004, 48 (2): 111-118.

    CAS  PubMed  Google Scholar 

  190. Afonina E, Neumann M, Pavlakis GN: Preferential binding of poly(A)-binding protein 1 to an inhibitory RNA element in the human immunodeficiency virus type 1 gag mRNA. J Biol Chem. 1997, 272 (4): 2307-2311. 10.1074/jbc.272.4.2307.

    CAS  PubMed  Google Scholar 

  191. Zolotukhin AS, Michalowski D, Bear J, Smulevitch SV, Traish AM, Peng R, Patton J, Shatsky IN, Felber BK: PSF acts through the human immunodeficiency virus type 1 mRNA instability elements to regulate virus expression. Molecular & Cellular Biology. 2003, 23 (18): 6618-6630. 10.1128/MCB.23.18.6618-6630.2003.

    CAS  Google Scholar 

  192. Wortman MJ, Krachmarov CP, Kim JH, Gordon RG, Chepenik LG, Brady JN, Gallia GL, Khalili K, Johnson EM: Interaction of HIV-1 Tat with Puralpha in nuclei of human glial cells: characterization of RNA-mediated protein-protein binding. J Cell Biochem. 2000, 77 (1): 65-74. 10.1002/(SICI)1097-4644(20000401)77:1<65::AID-JCB7>3.0.CO;2-U.

    CAS  PubMed  Google Scholar 

  193. Chepenik LG, Tretiakova AP, Krachmarov CP, Johnson EM, Khalili K: The single-stranded DNA binding protein, Pur-alpha, binds HIV-1 TAR RNA and activates HIV-1 transcription. Gene. 1998, 210 (1): 37-44. 10.1016/S0378-1119(98)00033-X.

    CAS  PubMed  Google Scholar 

  194. Krachmarov CP, Chepenik LG, Barr-Vagell S, Khalili K, Johnson EM: Activation of the JC virus Tat-responsive transcriptional control element by association of the Tat protein of human immunodeficiency virus 1 with cellular protein Pur alpha. Proc Natl Acad Sci U S A. 1996, 93 (24): 14112-14117. 10.1073/pnas.93.24.14112.

    PubMed Central  CAS  PubMed  Google Scholar 

  195. Cen S, Khorchid A, Javanbakht H, Gabor J, Stello T, Shiba K, Musier-Forsyth K, Kleiman L: Incorporation of lysyl-tRNA synthetase into human immunodeficiency virus type 1. J Virol. 2001, 75 (11): 5043-5048. 10.1128/JVI.75.11.5043-5048.2001.

    PubMed Central  CAS  PubMed  Google Scholar 

  196. Lama J, Trono D: Human immunodeficiency virus type 1 matrix protein interacts with cellular protein HO3. J Virol. 1998, 72 (2): 1671-1676.

    PubMed Central  CAS  PubMed  Google Scholar 

  197. Villace P, Marion RM, Ortin J: The composition of Staufen-containing RNA granules from human cells indicates their role in the regulated transport and translation of messenger RNAs. Nucleic Acids Res. 2004, 32 (8): 2411-2420. 10.1093/nar/gkh552.

    PubMed Central  CAS  PubMed  Google Scholar 

  198. Giles KE, Beemon KL: Retroviral splicing suppressor sequesters a 3' splice site in an ~50S aberrant splicing complex. Mol Cell Biol. 2005, 25: 4397-4405. 10.1128/MCB.25.11.4397-4405.2005.

    PubMed Central  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

A.W.C. is the recipient of a Scientist Award from The Ontario HIV-1 Treatment Network (OHTN) and his research is currently supported by grants from the Canadian Institutes of Health Research (CIHR, Grant #MOP-15103) and the OHTN (#ROGC114). M.T.M. is supported by the Public Health Service Grant from the National Cancer Institute, USA (Grant #R01 CA78709). A.J.M. is supported by a CIHR New Investigator Award and work in his laboratory is supported by grants from the CIHR (Grant #MOP-38111 & MOP-56974), the Canadian Foundation for Innovation (Project #6848) and the Canadian Foundation for HIV-1/AIDS Research (Grant #17724).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Andrew J Mouland.

Additional information

Competing interests

The author(s) declare that they have no competing interests.

Authors' contributions

All authors contributed equally to the inception and writing of the manuscript.

Authors’ original submitted files for images

Rights and permissions

Open Access This article is published under license to BioMed Central Ltd. This is an Open Access article is distributed under the terms of the Creative Commons Attribution License ( https://creativecommons.org/licenses/by/2.0 ), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Cite this article

Cochrane, A.W., McNally, M.T. & Mouland, A.J. The retrovirus RNA trafficking granule: from birth to maturity. Retrovirology 3, 18 (2006). https://doi.org/10.1186/1742-4690-3-18

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1742-4690-3-18

Keywords