Next Article in Journal
The Role of the Equine Herpesvirus Type 1 (EHV-1) US3-Encoded Protein Kinase in Actin Reorganization and Nuclear Egress
Previous Article in Journal
Morbillivirus Experimental Animal Models: Measles Virus Pathogenesis Insights from Canine Distemper Virus
Previous Article in Special Issue
Orchestrating the Selection and Packaging of Genomic RNA by Retroviruses: An Ensemble of Viral and Host Factors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Cross- and Co-Packaging of Retroviral RNAs and Their Consequences

1
Department of Microbiology & Immunology, College of Medicine and Health Sciences (CMHS), United Arab Emirates University (UAEU), P.O. Box 17666, Al Ain, United Arab Emirates
2
Zayed Bin Sultan Center for Health Sciences, CMHS, UAEU, P.O. Box 17666, Al Ain, United Arab Emirates
3
Department of Biochemistry, CMHS, UAEU, P.O. Box 17666, Al Ain, United Arab Emirates
*
Authors to whom correspondence should be addressed.
Submission received: 11 July 2016 / Revised: 3 October 2016 / Accepted: 3 October 2016 / Published: 11 October 2016
(This article belongs to the Special Issue RNA Packaging)

Abstract

:
Retroviruses belong to the family Retroviridae and are ribonucleoprotein (RNP) particles that contain a dimeric RNA genome. Retroviral particle assembly is a complex process, and how the virus is able to recognize and specifically capture the genomic RNA (gRNA) among millions of other cellular and spliced retroviral RNAs has been the subject of extensive investigation over the last two decades. The specificity towards RNA packaging requires higher order interactions of the retroviral gRNA with the structural Gag proteins. Moreover, several retroviruses have been shown to have the ability to cross-/co-package gRNA from other retroviruses, despite little sequence homology. This review will compare the determinants of gRNA encapsidation among different retroviruses, followed by an examination of our current understanding of the interaction between diverse viral genomes and heterologous proteins, leading to their cross-/co-packaging. Retroviruses are well-known serious animal and human pathogens, and such a cross-/co-packaging phenomenon could result in the generation of novel viral variants with unknown pathogenic potential. At the same time, however, an enhanced understanding of the molecular mechanisms involved in these specific interactions makes retroviruses an attractive target for anti-viral drugs, vaccines, and vectors for human gene therapy.

1. Introduction

Retroviruses are a special a category of viruses that contain two copies of unspliced full-length genomic RNA (gRNA) as their genome rather than DNA. Similar to eukaryotic cellular RNAs, retroviral gRNA is a 7-to-10-kb terminally redundant molecule that contains a 5’ guanosine cap and a 3’ polyadenosine (poly(A)) tail [1] that is packaged as an RNA dimer [2,3]. The efficient and specific packaging of two copies of plus strand full-length gRNA into the virus particle from a large pool of cellular and other viral RNAs in the cytoplasm by the assembling virion is considered one of the hallmarks of retroviral life cycle (reviewed in [3,4,5,6,7,8,9,10]). This process of specific encapsidation involves the recognition of particular sequences known as the packaging signal (psi, Ψ, or E) by the zinc finger domain of the nucleocapsid (NC) binding site of Gag polyprotein (reviewed in [3,5,6,8,10]). Following dimerization of two strands of RNA genome close to their 5’ ends, retroviral RNA genome is packaged as a non-covalently linked dimer; therefore, both the process of RNA dimerization and that of RNA packaging are thought to be closely interlinked ([2] and reviewed in [3,5,6,7,8,11,12,13,14,15,16,17,18,19,20]). Consistent with this, for almost all retroviruses, determinants of RNA packaging and dimerization map to the same 5’ end region (100–400 nucleotides; nts) of the gRNA (reviewed in [3,5,6,21]). The only exception seems to be in human foamy viruses (HFV), where RNA dimerization does not seem to be a prerequisite for packaging [22].
Although gRNA packaging is a critical step of the fidelity of retroviral particle assembly, it can lead to cross-/co-packaging of heterologous gRNAs expressed endogenously or exogenously following infection of the same cell by two different retroviruses. This is despite the fact no sequence conservation between the packaging signals of different retroviruses has been found. Rather, it has been shown that the packaging sequences of all known retroviruses assume a higher order structure comprising different structural motifs (reviewed in [3,5,6,8,23]). Thus, regardless of their primary sequence, these structural motifs have been strongly associated with retroviral RNA encapsidation and could explain the phenomenon of RNA cross- and co-packaging among diverse retroviruses [7,24,25,26,27,28,29,30,31]. This review summarizes the current understanding of the gRNA encapsidation determinants of a diverse group of retroviruses, elaborates on what is known about the interaction of these structural elements with their cognate structural Gag proteins, and then discusses the current understanding of cross-/co-packaging among these retroviruses. Cross-packaging refers to the ability of structural proteins of one virus to recognize and encapsidate the gRNA from another virus, while co-packaging refers to the ability of two RNAs from distinct retroviruses to be packaged together into one virus particle. A better understanding of the phenomena of cross-/co-packaging among retroviruses should provide not only insights into the characterization of cardinal packaging determinants both at the RNA and protein levels among retroviruses, but also the possibility of generating novel recombinant viruses with unknown pathogenic potential. Moreover, it should also help improve the design and safety of retroviral vectors for human gene therapy.

2. How Similar Is the Structural Organization of Packaging Elements within Various Retroviruses?

The packaging signals have been identified and characterized for a number of simple and complex retroviruses using deletion and substitution analyses. They are generally located between the primer binding site (PBS) and the initial part of the gag gene (reviewed in [5,6,32,33]). Figure 1 and Table 1 summarize the location of the primary packaging determinants of several genus of retroviruses, from the simple alpharetroviruses to the complex lentiviruses and the eccentric spumaviruses. As can be observed, the packaging determinants can either encompass the entire 5’ untranslated region (5’ UTR), as illustrated for human immunodeficiency virus type 1 (HIV-1), mouse mammary tumor virus (MMTV), bovine leukemia virus (BLV), and HFV, or part thereof, as is the case for human immunodeficiency virus type 2 (HIV-2), feline immunodeficiency virus (FIV), murine leukemia virus (MLV), and Rous sarcoma virus (RSV). For most retroviruses, the proximal region (100–400 nts) of the gag gene is important in this process. In some cases, other regions of the viral genome have also been implicated such as the pol in the case of HFV [34,35,36], env in the case of RSV [37], or the 3’ UTR in the case of avian leucosis viruses and MLV [38,39].
Furthermore, the packaging determinants can either be continuous or discontinuous at the primary sequence level with regions in between not playing any seminal role in gRNA packaging. The bipartite or multipartite nature of packaging determinants has been observed for several retroviruses, including FIV, Mason-Pfizer monkey virus (MPMV), BLV, and HFV (Figure 1 and Table 1). These observations suggest that gRNA packaging involves RNA-protein interactions at a structural level that brings together non-contiguous regions of the viral genome architecturally in one place to facilitate the process of gRNA packaging. The sequences within the 5’ UTR have been thought to contain cis-acting sequences important for protein binding (Gag) and dimerization initiation, while sequences in the gag gene have been speculated to stabilize the secondary structure of the region and facilitate the process of gRNA packaging into the assembling virions.
The first demonstration that cis-acting structural elements may be important for RNA packaging came with the observation that two hairpins were found to be involved in the packaging of the avian spleen necrosis virus (SNV) gRNA [58,62]. Interestingly, these structures could be replaced functionally with those from Moloney murine leukemia and murine sarcoma viruses with no effect on the replication of SNV-based vectors [58]. Thus, this was not only the first report of the importance of structural elements for RNA packaging, but also of cross-packaging among retroviruses. Since then, RNA folding algorithms, phylogenetic analyses, and biochemical probing have all revealed that the 5’ end of the retroviral genome folds into complex RNA secondary structures (stem loops) that affect retroviral gRNA packaging for retroviruses at biological levels. These structures contain helix-loop motifs with terminal purine-rich loops that specifically interact with the nucleocapsid protein to allow specific gRNA encapsidation into the viral particles.
The packaging determinants of HIV-1 and MLV have been the most extensively studied (reviewed in [5,8,33]). In HIV-1 and MLV, these regions are found predominantly downstream of the major splice donor (SD) in the 5’ UTR [54,55,63,64,65], and the gag-coding region [55,66,67]. However, in HIV-1, sequences upstream of the SD are also important [14,68]. The characteristic stem loops observed at the 5’ end of the HIV-1 genome correspond to important steps of the virus replication cycle that are either virus-specific (such as the trans-activation response element (TAR) loop important for transcriptional transactivation by the HIV-1 Tat protein), or more generic (such as the poly(A), U5/PBS, dimer initiation site (DIS), and psi stem loops) common to many retroviruses with critical roles both at the primary and structural levels. For example, both the sequence of DIS (CGGCCG and CUGCAG) and its location on a stem loop structure is important for gRNA dimerization in MPMV and MMTV, respectively [69,70]. The structural motifs at the 5’ end of the gRNA also interact at tertiary levels with each other and other parts of the viral genome for proper function. For example, the poly(A) stem loop of HIV-1 is involved in long-range interactions (LRIs) with sequences in the matrix (MA), forming a pseudoknot [71]. Similarly, the U5 stem loop sequences are involved in long-range interactions (LRI) with the stem loop containing the Gag AUG in many retroviruses, including HIV-1 and 2, SIV, FIV, MMTV, and MPMV [69,70,71,72,73,74,75,76,77,78,79], as well as other plus-strand RNA viruses with icosahedral capsids [80]. Mutations that destabilize these interactions affect several important steps in the retroviral replication cycle, including RNA packaging and dimerization [10,48,49,71,72,79,81,82].
Detailed RNA structural determinants of two important beta retroviruses, MMTV and MPMV, as well as lentiviruses such as HIV-1 and FIV have recently been reported as employing selective 2’-hydroxyl acylation analyzed by primer extension (SHAPE) followed by validation of predicted structural motifs for RNA packaging and dimerization using mutational analysis in a biologically relevant replication assay [69,70,74,76,79,82,83]. Emerging themes from these and other studies (reviewed in [3,8,33]) on RNA packaging can be summarized as follows:
1. The 5’ ends of retroviral genomes, starting from R to 100–400 nts into the gag gene fold into several stable, but flexible, helix-loop motifs termed stem loops (SLs) with purine-rich loops important for protein binding (Figure 1).
2. The RNA structure is held together by LRIs between the U5 and Gag sequences that occlude Gag AUG. Disruption of the LRIs affects both RNA dimerization and packaging.
3. Dimerization is thought to be a prerequisite for RNA packaging and is initiated by a GC-rich palindromic DIS containing a canonical GC dyad that is present in close proximity or overlapping the packaging determinants.
4. The DIS itself can be found in both “occluded” and “accessible” RNA conformations, allowing the cells to regulate ratios of monomeric and dimeric forms of RNA, the former of which is suited for translation (Gag AUG-accessible), while the latter for RNA packaging (Gag AUG-occluded).
5. Initiation of genomic RNA dimerization takes place via “kissing-loop” interactions through DIS on two RNA molecules (intermolecular interactions), which leads to the formation of the dimer linkage structure (DLS) that allows for the differentiation of gRNAs from mRNAs.
6. Furthermore, these conformational changes or “switches” result in the exposure of the otherwise hidden or obscured purine-rich nucleocapsid binding sites on the RNA (single-stranded (ss)purines-sspurines), initiating the process of Gag binding via the NC domain and multimerization on the RNA, leading to RNA encapsidation and virion assembly.
7. Depending upon the virus, the recruitment of dimerized gRNA by Gag takes place either in the nucleus (RSV and MLV), the pericentriolar region (MMTV and MPMV), the nuclear membrane (FIV), or the cytoplasm (HIV-1), followed by virion assembly either in the cytoplasm or the plasma membrane.
An interesting spatial feature observed among packaging signals of different retroviruses is the relative distance between the packaging signal and the 5’ methyl cap of the gRNA. This distance appears to be ~300–400 nt long at the 3’ end of the cap site for a number of simple and complex retroviruses, including SNV, RSV, MLV, MPMV, HIV-1, HIV-2, and SIV [84,85,86]. This distance is speculated to be critical for the encapsidation function and may relate to the competition between translation and encapsidation functions of the gRNA, which are necessary to ensure optimal replication efficiency of the virus [86]. How retroviruses maintain pools of gRNA to be used for Gag/Pol protein translation and as substrates for RNA packaging is not entirely clear, but the emerging picture suggests a role of internal ribosome binding sites (IRES) present within the 5’ UTR as well as within coding sequences of the genome that allows for the translation of Gag, independent of cap-dependent translation [87]. Having multiple mechanisms to ensure structural protein synthesis may be important for RNA packaging and virus assembly. This may also have relevance for the co-translational mechanism of gRNA packaging proposed for some retroviruses, where the primary packaging determinant is found upstream of the major splice donor and thus present on both genomic and spliced RNAs, such as RSV [59,88] and HIV-2 [42]. Accordingly, low amounts of Gag stimulate translation of Gag proteins from the full-length viral RNA in a cap-dependent manner. An increasing amount of Gag concentration in the cell then stimulates binding to the packaging signal at the 5’ UTR/beginning of gag in a co-translational manner [42,89,90,91]. Conformational changes then lead to multimerization of the protein on the gRNA scaffold, causing repression of translation from the cap structure while allowing protein synthesis from the IRES, thus stimulating RNA packaging and virus assembly [87].
Finally, the mechanism behind how viral genomic RNA is packaged preferentially to that of spliced viral RNAs in retroviruses where the primary packaging determinants are present in both spliced and unspliced RNAs is not entirely clear. It has been observed that, for HIV-1 and RSV, up to 10% of the packaged RNA may be spliced [59,92]. In the case of RSV, Env mRNA is the only spliced message of the virus translated on endoplasmic reticulum (ER)-associated ribosomes. Thus, the translated signal peptide sequesters the complex into the rough endoplasmic reticulum away from the Gag protein. The Gag protein, on the other hand, is produced on free cytoplasmic ribosomes, ready for binding to gRNA for virion assembly in a co-translational manner [89,90]. The spatial separation of the two mRNAs thus allows for selection for the genomic and not the spliced mRNA for packaging in RSV. Similarly, for HIV-2, specificity may be maintained by co-translational packaging of the Gag-encoding unspliced message [42,91], which also limits the amount of Gag available that could capture other spliced mRNAs [42].

3. What Determines the Specificity of gRNA Packaging among Retroviruses?

The packaging determinants of retroviral gRNA interact with the highly basic NC domain of the Gag precursor polyprotein to capture the gRNA for packaging into the assembling virus particles [3,5,6,8,11,21,32,83]. Elimination of the NC domain from Gag disrupts the specific packaging of HIV-1 gRNA, while packaging of cellular RNAs is supported by basic residues in the MA domain of Gag [1,93,94]. Depending upon the viral strain analyzed, NC contains either one to two evolutionarily conserved, but distinct, Cys-His boxes [95,96] that can sequester Zn2+ ions, allowing for high affinity NC-gRNA interactions [97,98,99,100,101]. Spumaviruses are an exception to this model in that that they contain arginine-rich motifs instead of the zinc fingers. The Cys-His boxes contain conserved CCHC arrays (C-X2-C-X4-H-X4-C where C = Cys, H = His, Xn = n number of amino acids) that are variable between different retroviruses (Figure 2). Flanking (and sometimes overlapping) the Cys-His boxes are stretches of basic amino acids important for specific RNA binding of NC [102] that are found in many retroviruses. Mutations in the conserved CCHC boxes can significantly reduce genomic RNA packaging and give rise to non-infectious virions [65,102,103,104,105,106,107,108,109,110,111,112,113,114,115]. Mutations in the flanking regions of Cys-His boxes, which contain basic residues, also result in severe replication defects [102,116,117,118]. Furthermore, studies with chimeric Gag proteins have revealed that specificity of RNA packaging could be interchanged when the NC domains between two viruses were switched, such as those between MLV and RSV [119] and between MLV and HIV-1 [110,120]. Since the amino acid sequence of the regions flanking the Cys-His boxes are quite diverse (see Figure 2), the flanking regions of the NCs were shown to be “functionally conserved” by replacing a flanking region from SNV with that of MLV and vice versa, which did not affect either Gag polyprotein or NC function for virus RNA encapsidation or virion assembly [121]. These studies imply that NC and its flanking regions are a critical component determining specificity of gRNA packaging, though they may not be the sole determinant (see below).
The zinc finger domains of HIV-1 NC bind specifically to the viral genomic RNA via the encapsidation signal [122]. The two zinc fingers are not functionally equivalent or interchangeable as evidenced by experiments in which the two motifs were either switched or duplicated [108]: the first Cys-His box plays a more prominent role in encapsidation and is more selective for genomic viral RNA for encapsidation. This is evidenced by the fact that mutations in the first Cys-His box have a more drastic effect on gRNA packaging (100-fold) than those in the second box (10–30-fold) [105,106]. Extensive structural analysis of NC-psi sequences has been carried out using nuclear magnetic resonance (NMR) studies defining the actual molecular interactions that take place between NC and the packaging signals (reviewed in [8]). NC probably binds to dimeric genomes, and this interaction leads to the multimerization of the Gag polyproteins, leading to virus assembly [123,124].
Despite the fact that NC has been shown to be important for gRNA packaging, several lines of evidence suggest that NC may not be the only determinant of specific gRNA packaging, and other Gag domains may also be involved. For one, NC is a highly basic protein; as such, it has non-specific RNA binding activity [120]. Furthermore, RNA packaging experiments using chimeric Gag proteins between HIV-1 and MMTV where the NC domains of the respective viruses were swapped revealed that these viruses still packaged their cognate gRNAs preferentially [125], unlike what was observed for chimeras between MLV and RSV and between MLV and HIV-1 discussed above. Moreover, several independent studies have implicated other regions of the Gag polyprotein, including MA [10], capsid (CA) [126], the p2 spacer peptide between CA and NC [40,127,128], and the terminal p6 late domain [110]. Considering that the Gag polyprotein is the initial recruiter of gRNA and not the cleaved NC protein, this suggests that specific selection of gRNA from other cellular and spliced RNAs is more complex and happens in the context of the whole polyprotein, as has recently been shown for HIV-1 [83].
The intracellular sites of initial interaction between gRNA and NC domain of the Gag polyprotein also have important implications for virion RNA packaging and assembly. These interactions can occur in the nucleus or the cytoplasm, depending upon the nature of the retrovirus [33]. The simple retroviruses such as RSV and MLV use the help of the Gag proteins containing nuclear localization signals and the cellular importin-α/β proteins to enter the nucleus to export the gRNA to the cytoplasm [129]. In these viruses, the gRNA is thought to dimerize in the nucleus before interactions of the packaging signal with the NC domain of the viral Gag polyprotein [130]. This gRNA/Gag complex is then transported into the cytoplasm via the cellular chromosome region maintenance 1 (CRM1) RNA export pathway for the eventual virus assembly at the plasma membrane [131]. The complex retroviruses such as HIV-1 and FIV, on the other hand, can export the genomic unspliced RNA using their accessory genes (rev) involved in the nucleocytoplasmic passage of unspliced, full-length gRNA, which would otherwise get spliced by the cellular splicing machinery [132]. Thus, for these viruses, the gRNA exits the nucleus and first interacts with the Gag proteins in the cytoplasm, either at the nuclear membrane, as is the case with FIV, or the cytosol, as is the case with HIV [133,134]. Either way, the gRNA/Gag polyprotein complexes are then directed to the cell membrane for final assembly into the virus particles. This process employs the matrix domain of the Gag polyprotein that not only has the N-myristoyl moiety for plasma membrane targeting, but also a stretch of basic amino acids that can interact with both gRNA and plasma membrane-specific phospholipids for proper targeting [135,136].

4. How Do Retroviruses Cross-Package Each Other’s Genomes?

Based on the evidence presented above, it is clear that the packaging determinants of retroviruses at the three-dimensional RNA structural level are spread out and elastic, composed of primary, secondary, as well as tertiary interactions. This lack of rigidity and absolute specificity is necessary for the numerous conformational changes that need to take place to allow successful packaging of the gRNA preferentially over the spliced as well as the cellular mRNAs. A direct consequence of this flexibility is RNA pseudotyping or cross-packaging; i.e., packaging of an RNA by heterologous viral particles that can take place among evolutionarily related yet molecularly different retroviruses [24,25,28,29,30,31,40,58,137,138,139,140,141]. Furthermore, co-packaging of heterodimeric RNAs from two divergent retroviruses can also take place [26,27,31], resulting in recombinant viruses with unknown pathogenic potential, if successful reverse transcription and template switching can take place following the co-packaging of two genetically distinct RNAs.
Thus, cross- and co-packaging among retroviruses has been investigated to determine the potential hazards and safety of using retroviral-based vectors for human gene therapy since patients being treated could be infected by other retroviruses, while the human genome itself contains endogenous retroviruses that could provide substrates for cross- or co-packaging [142,143,144,145,146,147]. Furthermore, a better understanding of these phenomena should help in the development of novel retroviral vectors (both recombinant/hybrid and otherwise) for safer and more efficient gene delivery systems into humans [148,149,150,151,152,153,154]. Towards this end, genomes of several different retroviruses from different hosts have been explored, including the simian lentivirus SIV [30,137,155] the feline lentivirus FIV [156,157], the avian retrovirus SNV [28], the mouse retroviruses MLV and MMTV [158,159], and even the more simian-like human lentivirus, HIV-2 [137]. These studies have shown that it is possible to develop novel hybrid vector systems (e.g., SNV/HIV-1 [28] and SIV/HIV-2 [137]) that maintain high transduction efficiencies with possibly reduced pathogenic potential.
One of the earliest examples of cross-packaging among retroviruses was observed between the avian SNV and murine MLV retroviruses when it was shown in a series of elegant studies that a SNV helper cell line producing SNV structural proteins could cross- and co-package RNA from the distantly related retrovirus, MLV, and generate recombinants [31,58,141,160]. Since then, a number of human and animal retroviruses have been studied for their cross-packaging potential using genetic complementation and/or biological replication assays. For example, among complex retroviruses, several studies have revealed that lentiviruses such as HIV-1, SIV, and even the more distantly related FIV, can cross-package each other’s genomes [29,30,161]. Subsequently, these observations were expanded to the beta retroviruses, when reciprocal cross-packaging between the two distantly related type B and D retroviruses—MMTV and MPMV respectively—was reported [25]. Not only that, reciprocal cross-packaging has also been demonstrated between HIV-1 and SIV and between MPMV RNA and proteins [24]. In most of these cases, cross-packaging was demonstrated to be primarily due to the recognition of the packaging signal by the heterologous proteins.
In addition to reciprocal cross-packaging, non-reciprocal packaging has also been observed, between simple retroviruses where the SNV genome could not be cross-packaged by MLV proteins [58,138,141,160]. Moreover, RNA from a complex retrovirus (HIV-1) could be cross-packaged by proteins from a simple retrovirus, SNV, but not vice versa [28]. On the contrary, RNA from a complex retrovirus (FIV) could not be cross-packaged by proteins from a simple retrovirus (MPMV), while FIV proteins were efficient at cross-packaging MPMV RNA [161].
Non-reciprocal cross-packaging has also been observed even among more closely related retroviruses. For example, the human lentivirus HIV-2 that can also cause acquired immunodeficiency syndrome (AIDS) like HIV-1, although to a much more restricted extent, is phylogenetically closer to SIVs than to HIV-1. Interestingly, HIV-2 RNA has been shown to be cross-packaged by HIV-1 proteins, but HIV-2 proteins were unable to encapsidate and transduce HIV-1 RNA [40,137]. However, HIV-2 Gag chimeras that contained the HIV-1 NC and p2 domains were able to encapsidate HIV-1 RNA. HIV-2 Gag chimeras with only the HIV-1 NC domain also exhibited the ability to encapsidate HIV-1 RNA, although at a lower level than when both NC and p2 were present, suggesting that amino acids within the p2 domain also contribute to RNA packaging. The HIV-1 and HIV-2 NC proteins are quite similar with 60% homology at the amino acid level and only conservative amino acid difference in other parts of the protein, while the p2 domain is much less similar with only 35% homology at the amino acid levels. This supports the hypothesis that selective recognition may be spread over other regions of the Gag precursor [40,137]. One of the reasons for the observed packaging restriction could be the co-translational mode of packaging observed in HIV-2, which is distinct from the trans-packaging observed in HIV-1 [91]. Replacement of HIV-2 NC with that of HIV-1 could have relieved that inhibition to allow cross-packaging perhaps by using the co-translational mode of packaging.
Table 2 summarizes many of the studies available in literature regarding cross-packaging of retroviral genomes and their subsequent transduction (propagation) into target cells. As can be seen, with the exception of HIV-2, the other lentiviral proteins tested (HIV-1, SIV, and FIV) were the most promiscuous for allowing RNA cross-packaging, as they could not only cross-package each other’s genomes, but also genomes from MPMV, MMTV, and MLV (tested only for HIV-1). On the other hand, the HIV-2 and MLV proteins seemed the least promiscuous, unable to cross-package RNA from any of the tested viruses. HIV-2 could neither package HIV-1 nor SIV RNAs despite close genetic relationship with these viruses. Similarly, MLV proteins could neither package SNV RNA (an avian virus that is closer to the mouse retroviruses than other avian viruses), or RNAs from other more distantly related retroviruses such as HIV-1 or MMTV. On the contrary, SNV proteins could cross-package RNAs from both MLV and the distantly related HIV-1.
Interestingly, even though HIV-1 proteins were the most promiscuous among the retroviruses tested, they were unable to cross-package SNV RNA. Among the beta retroviruses, both MPMV and MMTV proteins could cross-package RNAs from the tested heterologous viral RNAs with the exception of FIV RNA that could not be cross-packaged by MPMV proteins. This is despite the fact MPMV could cross-package both HIV-1 and SIV RNAs. The inability of SNV RNA to be cross-packaged by either HIV-1 or chimeric MLV NC proteins or the FIV RNA by MPMV proteins could probably be explained by restrictions perhaps not at the RNA end, but at the protein level, such as the NC and matrix proteins as has been suggested earlier for some retroviruses [110,125]. Finally, the cross-packaging of MLV RNA has also been observed by both RSV and BLV proteins in addition to HIV-1, even though phylogenetically MLV RNA is far from these viruses.
Thus, the ability to cross- or co-package a retroviral genomic RNA into heterologous structural proteins is a lot more complex and involved. It depends upon not only the packaging determinants on the gRNA and the zinc finger domains of the NC protein, but also by the type of viral RNA available for packaging, its cellular location (nuclear versus cytoplasmic), mechanistic differences in packaging (cis versus trans), the role of Gag domains other than NC, etc. Considering that gRNA in the virion does not exist as a linear, dimeric species but rather in a highly compact and condensed form as a nucleoprotein complex [162] adds another layer of complexity to this phenomena, making empirical testing still the best way to determine whether two viral genomes can successfully cross- or co-package, or both, their gRNAs into a particular virus particle.

5. What Determines the Propagation Capabilities of the Cross-Packaged RNAs?

Table 2 also summarizes the capability of different retroviruses to propagate (transduce) the cross-packaged RNAs by the heterologous viral proteins. As can be seen, the propagation of the cross-packaged RNA into the target cells has been found to be quite frequent among retroviruses, though to a lesser extent than RNA cross-packaging. This suggests that the successful propagation of the cross-packaged RNA must involve compatible interactions between the cis- and trans-acting sequences needed for reverse transcription and integration. This includes the ability of the reverse transcriptase enzyme of the packaging viral particles to recognize the PBS and polypurine tracts (PPT), the viral integrase to recognize the attachment (att) sites of the cross-packaged RNA, or both.
Overall, Table 2 shows that, if a viral RNA could be cross-packaged, nearly 60% of the time it could also successfully propagate and transduce the target cells. However, there were exceptions. For example, the most promiscuous lentiviruses, HIV-1, SIV, and FIV, could cross-package each other’s RNAs and propagate them successfully; they could also cross-package MPMV (a non-lentivirus) RNA as well, but all three lentiviruses were unable to propagate the cross-packaged MPMV RNA into the target cells [24,161]. Similarly, MPMV proteins could cross-package HIV-1 and SIV RNAs (though not FIV), but could not propagate these RNAs further, pointing towards the incompatibility of the MPMV reverse transcriptase and integrase enzymes with the cis-acting sequences of the heterologous (HIV-1 and SIV) viral RNAs. A comparison of the cis-acting sequences important for viral RNA propagation among HIV-1, SIV, FIV, and MPMV revealed size differences in the length of PBS and PPT (shorter for MPMV than the lentiviruses) as well as the least homology of the MPMV att sites compared to HIV-1, SIV, or FIV (Figure 3). These observations suggest that incompatibility between FIV reverse transcriptase and integrase enzymes with that of MPMV PBS, PPT, and att sites is the likely reason for the inefficient RNA propagation observed between MPMV and the lentiviruses. Similar observations have been reported when HIV-1 att sites were substituted by those from FIV in HIV-1 RNA, resulting in only a one-third reduction in integration efficiency, while substitution with the more distantly related MLV att sites led to nearly a complete abrogation of integration of the hybrid HIV-1 RNA [164].
As can be further observed from Table 2, HIV-2 and MLV were the two viruses that were the most restrictive in their ability to cross-package heterologous RNAs. SNV, on the other hand, was more like HIV-1, SIV, and FIV, and could not only cross-package HIV-1 and MLV RNAs, but also successfully propagate these RNAs into the target cells, suggesting compatibility of its reverse transcriptase and integrase enzymes for the cis-acting sequences of the heterologous viruses.
As mentioned earlier, the two type B and D betaretroviruses—MMTV and MPMV—could reciprocally cross-package their genomes; however, both viral proteins were unable to transduce the infected cells with each other’s cross-packaged RNAs [25]. Sequence analysis of the cis-acting elements of MMTV and MPMV RNAs revealed that MMTV PBS had a 72% sequence homology to MPMV, while MMTV PPT showed a 90% homology (Figure 3). However, comparison of att sites indicated a sequence homology of less than 50% (the U3 att was 40%, whereas the U5 att was 45%), suggesting problems with reverse transcription and integration. Thus, the sequences of RNase H and integrase enzymes were also compared, which were found to be only 32% and 49% homologous, respectively [25]. Based on these analyses and the fact that both of these viruses utilize tRNALys primers (Lys-1 and -2 for MPMV and Lys-3 for MMTV), it was proposed that these viruses probably could not transduce the cross-packaged RNAs either due to inadequate reverse transcription or integration of fully reverse-transcribed DNA, or both, resulting in failed propagation [25]. Several studies support these assertions [24,28,161,164,165]. In a study by Parveen et al. [28], the role of integrase in improving transduction efficiency of cross-packaged HIV-1 RNAs by SNV proteins was studied in particular. This was achieved by making chimeric Gag/Pol packaging constructs where different domains of SNV integrase were replaced by that of HIV-1. Only one of the chimeric integrase (IN) proteins was able to transduce the HIV-1 vector RNA, albeit at the same efficiency as the wild-type SNV packaging construct [28]. This study suggests that efficient propagation of cross-packaged RNAs requires overcoming multiple post-entry blocks of the virus replication cycle.

6. Packaging of Non-Viral RNAs into Retroviral Particles

How can one be assured that, once cross-packaging is observed, it is due indeed to the specific recognition of the packaging determinants and not due to either the known non-specific nucleic acid binding activity of NC [21] or other artifacts of the experimental system such as overexpression of the expressed RNA to be packaged into the cell? One way to demonstrate specificity of a newly mapped packaging signal is to clone the packaging signal in question onto a non-viral RNA and determine if that RNA can then be packaged specifically by the homologous protein. This approach has been used to confirm specificity of RNA packaging for a number of retroviruses, including MLV [55], BLV [51], and FIV [166]. The same approach was used to demonstrate specificity of cross-packaging between MPMV and MMTV [25], and that between HIV-1, SIV, and MPMV [24,25] as well. Thus, the cloning of the putative packaging determinants of MPMV and MMTV into pcDNA3-based vector RNAs confirmed that these non-viral RNAs could be co- and cross-packaged by the homologous and heterologous proteins [24,25]. To ensure that the results obtained were not due to overexpression of the vector RNAs, control vector RNAs using the same RNAs were expressed, but in the absence of the packaging signals. Together, these results reveal that neither the non-specific binding affinity of NC for RNA nor overexpression of the vectors is responsible for the cross-packaging activities being observed.

7. Packaging of Cellular RNAs into Retroviral Particles

It has been estimated that gRNA dimer is the dominant RNA species found in a virus particle accounting for ~50% [21] of the virion RNA content by weight along with 10–50 tRNA molecules used by the virus to prime reverse transcription [1]. The question that naturally arises is: what other RNAs are packaged into the virus particles, and are they packaged specifically? It has been shown that, other than gRNA, other cellular [21,167,168,169,170,171,172,173] and occasionally spliced viral RNAs [59,92,174,175,176] are also packaged. Packaging of the spliced viral mRNAs is no surprise since, depending upon the virus, the primary or secondary packaging determinants may be found upstream of the SD site; hence, these signals could be present in both the spliced and unspliced viral mRNAs, allowing their specific or semi-specific packaging into the virus particles. However, the role that these spliced viral mRNAs might play in the virus life cycle is not clear.
What about the cellular RNAs and their packaging into the virus particles? Cellular RNAs represent the remaining 50% of the virion RNA content in the same proportion as observed in the cells [1,177]. Since most of the cellular RNAs are composed of non-coding RNAs (ncRNAs), these are the types of RNAs also found in virions (reviewed in [178]). Among these, some have been observed to be enriched in virus particles such as the 7SL RNA, 5S rRNA, the spliceosomal U6 small nuclear RNA (snRNA), cytoplasmic Y RNA, and some cellular tRNAs [178]. On the contrary, no specific cellular mRNA has been found to be enriched in the virus particles. The mechanism of how the ncRNAs may be encapsidated into the virus particles is still under investigation and may require both specific and non-specific interactions with regions within Gag, Pol, or even the gRNA itself [178]. The discovery that some non-coding RNAs may be specifically packaged by the virions suggests novel ways of how retroviruses might use ncRNAs in their assembly and infectivity processes.

8. Co-packaging and Consequences of Co-Packaging among Diverse Retroviruses

All retroviruses encapsidate two copies of full-length gRNAs as a non-covalently linked dimer into the virus particles. Co-packaging is thus a natural part of the retroviral life cycle during which retroviruses package their gRNAs in pairs, even when the cells have limiting amounts of gRNA [179]. The norm is that the two gRNA copies are identical when a cell is infected with one type of virus, resulting in homodimers. However, if a cell is infected with more than one virus, or contains endogenous viruses that express viral RNA with compatible DIS, a heterodimer may form that could then be packaged by the expressed Gag proteins into retroviral particles. Co-infection with divergent retroviruses or co-transfection of experimentally produced retroviral vectors have been shown to result in the co-packaging of two different retroviral RNAs within a virus particle. For example, HIV-1 RNA can heterodimerize with either MLV or RSV RNAs [180], while MLV can co-package SNV [31]. A direct consequence of co-packaging is the exchange of genetic information during reverse transcription resulting in the formation of retroviral recombinants. For instance, the co-packaging and exchange of genetic information between genetically different retroviruses such as SNV and MLV [31] as well as between HIV-1 and HIV-2 have been reported [26]. The generation of viral variants can provide evolutionary advantages to the recombinants such as the acquisition of new or unknown pathogenic potential that may cause diseases, allow escape from the host’s immune system, or provide the recombinants with expanded host tropism [179,181,182].

8.1. Role of Dimerization in Co-Packaging

RNA dimerization allows for the formation of a unique RNA configuration—the dimer linkage structure—that helps distinguish gRNAs from mRNAs [3,179]. Other than ensuring specificity of gRNA packaging, dimerization is an essential prerequisite for the cross- and co-packaging of gRNAs. Packaging of dimeric RNAs is thought to increase the fidelity of genomic RNA replication, allowing retroviruses to generate full-length intact proviruses, even though persistent nicking of gRNA is observed [3,179], while co-packaging allows for virus evolution (increased genetic diversity) due to the generation of new recombinant proviruses. As discussed earlier, the efficiency of heterodimer formation is affected by the cellular site and timing of gRNA dimerization which has been shown to be different in different retroviruses (reviewed in [3,33]). As a result, different types of dimers are possible: viruses that dimerize late in the replication cycle such as in the cytoplasm (e.g., HIV-1 and FIV) have the possibility to form both homodimers as well as heterodimers, whereas viruses undergoing dimerization soon after transcription in the nucleus (such as MLV and RSV) primarily form homodimers even if other compatible RNAs are present. Thus, viruses such as HIV-1 that dimerize in the cytoplasm have a higher potential to co-package gRNAs as heterodimers [179,181].
The co-packaging frequency of the heterologous RNA can be influenced by their DIS sequences. When the two DIS sequences of HIV-1 and HIV-2 were mutated to simultaneously discourage RNA homodimerization and encourage RNA heterodimerization, HIV-1 and HIV-2 RNAs could heterodimerize using the DIS sequences prior to packaging [183,184]. Hence, it seems that the major determinant for co-packaging is to have a complementary DIS sequence among the two RNAs. Studies have shown that RNAs carrying palindromic or non-palindromic DIS sequence can heterodimerize if the DIS from two distinct viruses are complementary to each other [4,7,27,185,186]. Consistent with this, in vitro-transcribed trans-complementary mutants maintaining the central “GC” dyads have shown efficient RNA dimerization, authenticating that an intermolecular interaction mediated by the DIS is required for the formation of a stable RNA dimer [19,69].

8.2. Recombination Potential of Retroviruses

A salient characteristic of retrovirus replication is the high prevalence of genetic recombination that can assort mutations in the viral genome to increase replication fidelity and virus diversity. Recombinants are often generated after co-infection of a cell by two genetically distinct viruses, a feature first observed in avian tumor viruses [187] and later in other retroviruses as well [188,189,190]. The presence of multiple receptors on a target cell facilitates dual infection which is often cell-mediated [191,192,193]. The two distinct genomes need to be co-packaged, leading to heterozygous virion formation, for successful recombination to occur [194,195].
Recombination is a consequence of reverse transcription, a process that requires the reverse transcriptase (RT) to jump several times between the two genomic RNA templates to create a single DNA copy of the genome. It has been estimated that, during the process of reverse transcription, RT dissociates at least eight times for a 10-kb viral genome [196,197,198]. The template switching can be intermolecular [195,199,200,201], resulting in recombination [199,202,203,204,205]. It can also be intramolecular, resulting in deletions, deletions with insertions, or insertions and duplications [196,206,207], or it can be in a fashion that the first strand transfer step in reverse transcription is an intermolecular event while the second transfer is intramolecular [208]. Thus, generation of a recombinant virus requires multiple steps to happen successfully, such as (1) dual infection of a cell; (2) expression of both viral genomes; (3) dimerization of the heterologous RNAs; (4) successful co-packaging of the heterodimeric gRNAs; (4) RT template switching between the heteromeric co-packaged RNAs; and (5) integration and expression of the recombinant DNA. Despite this complexity, retroviruses still possess one of the highest recombination rates among viruses [181].
A comparison of sequences among different retroviruses suggests that ancient recombination events between different retroviruses may have occurred [2]. A classic example of this is the avian retrovirus SNV. The gag/pol region of SNV is related to that of MLV, whereas the env region is related to that of the simian virus MPMV [139]. This assertion is further supported by the observation that SNV and MPMV Env proteins bind to the same receptor during infection [139,209]. Another evidence for genetic exchange and recombination has been observed with the squirrel monkey retrovirus (SMRV), a type D retrovirus in which the env sequences were found to be homologous to the type C baboon endogenous retrovirus (BaEV), while the pol gene to type A, type B and avian type C retroviruses [210]. The LTRs sequences of SMRV were found to be present in another New World species: the skunk. However, the PBS was unique to type D retrovirus [210]. These kind of recombination events are likely to occur more than once in the evolution of retroviruses and therefore retroviruses are believed to have originated as a consequence of enormous genetic exchanges, following multiple contact or infections of the host with different species of retroviruses [211].
What are some of the elements that affect recombination frequency? Longer lengths of homology are one of the driving forces that influence a successful template switch [181]. Thus, recombinants between highly similar strains are generated at the highest frequencies than more diverse viruses [212,213]. Differences in site of dimerization of the two viral heterologous RNAs can also contribute to differences in recombination frequencies. For example, even though the RTs of HIV-1, SNV, and MLV appear to have similar processivity, the recombination rates of gammaretroviruses are 10-fold lower than those of lentiviruses [214,215]. This has now been attributed to the early commitment to dimerization of gammaretroviruses RNAs while still within the nucleus, and results in favoring homo- rather than heterodimers [179]. HIV-1 exports its gRNA into the cytoplasm before dimerization can take place. This later stage of commitment has been shown to result in the random distribution of heterodimers, which naturally result in higher frequencies of recombination. Similarly, the export pathway selected by the transcribed viral RNAs to exit the nucleus can affect its co-packaging [7]. Thus, two viruses that use the same export pathway for their RNA transport (e.g. nuclear RNA export factor 1 (NXF1) pathway) are also co-packaged more efficiently [7].
The secondary RNA structure of the gRNA also influences the recombination efficiency. Specific hairpin structures on the template RNA have been shown to stimulate RT for strand transfer to the acceptor DNA in the presence of NC protein [216]. In fact, different groups have identified recombination hotspots associated with RNA hairpins in the R region [217], gag [218], pol [219], and env [220,221]. Given the association between RNA stem-loops and recombination, these identified structures may facilitate recombination and regulate translation as well [181]. Similarly, the NC protein affects recombination efficiency by influencing various steps of viral replication. In addition to the already-discussed roles of NC in specific gRNA packaging and virus assembly, NC also affects tRNA placement onto the template RNA, dimerization, reverse transcription, and integration of the provirus [21,95,96,103,110,117,120,125,222,223,224,225,226].

9. Concluding Remarks

The ability of heterologous retroviruses to cross- and co-package their gRNAs once again reinforces the notion that retroviruses are highly promiscuous genetic mobile elements. This suggests that, if co-infected, co-packaged, and heterodimerized, these viruses could provide an enormous substrate for genetic recombination that might lead to the generation of replication-competent viral variants with unknown pathogenic potential, as well as drug- and vaccine-resistant viral variants. Therefore, delineating the requirements and factors that influence heterologous RNA cross- and co-packaging should help better understand the phenomenon of not only the packaging, but also the evolution of variant viruses with unknown pathogenic potential. More importantly, they should also allow improvements in the design of currently used retroviral vectors for human gene therapy as well as other gene-based approaches to treatment so as to minimize the chances of recombination, which is currently considered the most significant setback in using retroviral vectors in clinical trials.

Acknowledgments

The authors would like to thank the editorial assistance of Akhil Chameettachal and Fathima Nuzra Nagoor Pitchai for reading the manuscript and providing their input. Research in F.M. and T.A.R. research laboratories was funded by UAEU Program for Advanced Research (UPAR) and center-based interdisciplinary grants from the Office of the Deputy Vice Chancellor of Research and Graduate Studies at the UAE University and the Terry Fox Fund for Cancer Research to T.A.R.

Author Contributions

L.M.A., T.A.R., and F.M. all helped to develop and write the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rulli, S.J.; Hibbert, C.S.; Mirro, J.; Pederson, T.; Biswal, S.; Rein, A. Selective and Nonselective Packaging of Cellular RNAs in Retrovirus Particles. J. Virol. 2007, 81, 6623–6631. [Google Scholar] [CrossRef] [PubMed]
  2. Coffin, J.M.; Hughes, S.H.; Varmus, H.E. (Eds.) Retroviruses; Cold Spring Harbor Laboratory Press: Cold Spring Harbor, NY, USA, 1997.
  3. Johnson, S.F.; Telesnitsky, A. Retroviral RNA Dimerization and Packaging: The What, How, When, Where, and Why. PLoS Pathog. 2010, 6, e1001007. [Google Scholar] [CrossRef] [PubMed]
  4. Hussein, I.T.M.; Ni, N.; Galli, A.; Chen, J.; Moore, M.D.; Hu, W.-S. Delineation of the preferences and requirements of the human immunodeficiency virus type 1 dimerization initiation signal by using an in vivo cell-based selection approach. J. Virol. 2010, 84, 6866–6875. [Google Scholar] [CrossRef] [PubMed]
  5. D’Souza, V.; Summers, M.F. How retroviruses select their genomes. Nat. Rev. Microbiol. 2005, 3, 643–655. [Google Scholar] [CrossRef] [PubMed]
  6. Lever, A.M.L. HIV-1 RNA packaging. Adv. Pharmacol. 2007, 55, 1–32. [Google Scholar] [PubMed]
  7. Moore, M.D.; Nikolaitchik, O.A.; Chen, J.; Hammarskjöld, M.-L.; Rekosh, D.; Hu, W.-S. Probing the HIV-1 Genomic RNA Trafficking Pathway and Dimerization by Genetic Recombination and Single Virion Analyses. PLoS Pathog. 2009, 5, e1000627. [Google Scholar] [CrossRef] [PubMed]
  8. Miyazaki, Y.; Miyake, A.; Nomaguchi, M.; Adachi, A. Structural dynamics of retroviral genome and the packaging. Front. Microbiol. 2011, 2, 264. [Google Scholar] [CrossRef] [PubMed]
  9. Russell, R.S.; Liang, C.; Wainberg, M.A. Is HIV-1 RNA dimerization a prerequisite for packaging? Yes, no, probably? Retrovirology 2004, 1, 23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Lu, K.; Heng, X.; Summers, M.F. Structural determinants and mechanism of HIV-1 genome packaging. J. Mol. Biol. 2011, 410, 609–633. [Google Scholar] [CrossRef] [PubMed]
  11. Hibbert, C.S.; Mirro, J.; Rein, A. mRNA molecules containing murine leukemia virus packaging signals are encapsidated as dimers. J. Virol. 2004, 78, 10927–10938. [Google Scholar] [CrossRef] [PubMed]
  12. Paillart, J.-C.; Dettenhofer, M.; Yu, X.-F.; Ehresmann, C.; Ehresmann, B.; Marquet, R. First snapshots of the HIV-1 RNA structure in infected cells and in virions. J. Biol. Chem. 2004, 279, 48397–48403. [Google Scholar] [CrossRef] [PubMed]
  13. Baig, T.T.; Lanchy, J.-M.; Lodmell, J.S. HIV-2 RNA dimerization is regulated by intramolecular interactions in vitro. RNA 2007, 13, 1341–1354. [Google Scholar] [CrossRef] [PubMed]
  14. Berkhout, B.; van Wamel, J.L. Role of the DIS hairpin in replication of human immunodeficiency virus type 1. J. Virol. 1996, 70, 6723–6732. [Google Scholar] [PubMed]
  15. Lanchy, J.-M.; Ivanovitch, J.D.; Lodmell, J.S. A structural linkage between the dimerization and encapsidation signals in HIV-2 leader RNA. RNA 2003, 9, 1007–1018. [Google Scholar] [CrossRef] [PubMed]
  16. Lanchy, J.-M.; Lodmell, J.S. An extended stem-loop 1 is necessary for human immunodeficiency virus type 2 replication and affects genomic RNA encapsidation. J. Virol. 2007, 81, 3285–3292. [Google Scholar] [CrossRef] [PubMed]
  17. Laughrea, M.; Jetté, L. HIV-1 Genome Dimerization:  Kissing-Loop Hairpin Dictates Whether Nucleotides Downstream of the 5‘ Splice Junction Contribute to Loose and Tight Dimerization of Human Immunodeficiency Virus RNA. Biochemistry 1997, 36, 9501–9508. [Google Scholar] [CrossRef] [PubMed]
  18. McBride, M.S.; Panganiban, A.T. Position dependence of functional hairpins important for human immunodeficiency virus type 1 RNA encapsidation in vivo. J. Virol. 1997, 71, 2050–2058. [Google Scholar] [PubMed]
  19. Paillart, J.C.; Skripkin, E.; Ehresmann, B.; Ehresmann, C.; Marquet, R. A loop-loop “kissing” complex is the essential part of the dimer linkage of genomic HIV-1 RNA. Proc. Natl. Acad. Sci. USA 1996, 93, 5572–5577. [Google Scholar] [CrossRef] [PubMed]
  20. Paillart, J.; Marquet, R.; Skripkin, E.; Ehresmann, C.; Ehresmann, B. Dimerization of retroviral genomic RNAs: structural and functional implications. Biochimie 1996, 78, 639–653. [Google Scholar] [CrossRef]
  21. Berkowitz, R.; Fisher, J.; Goff, S.P. RNA packaging. Curr. Top. Microbiol. Immunol. 1996, 214, 177–218. [Google Scholar]
  22. Cain, D.; Erlwein, O.; Grigg, A.; Russell, R.A.; McClure, M.O. Palindromic sequence plays a critical role in human foamy virus dimerization. J. Virol. 2001, 75, 3731–3739. [Google Scholar] [CrossRef] [PubMed]
  23. Lu, K.; Heng, X.; Garyu, L.; Monti, S.; Garcia, E.L.; Kharytonchyk, S.; Dorjsuren, B.; Kulandaivel, G.; Jones, S.; Hiremath, A.; et al. NMR detection of structures in the HIV-1 5’-leader RNA that regulate genome packaging. Science 2011, 334, 242–245. [Google Scholar] [CrossRef] [PubMed]
  24. Al Shamsi, I.R.; Al Dhaheri, N.S.; Phillip, P.S.; Mustafa, F.; Rizvi, T.A. Reciprocal cross-packaging of primate lentiviral (HIV-1 and SIV) RNAs by heterologous non-lentiviral MPMV proteins. Virus Res. 2011, 155, 352–357. [Google Scholar] [CrossRef] [PubMed]
  25. Al Dhaheri, N.S.; Phillip, P.S.; Ghazawi, A.; Ali, J.; Beebi, E.; Jaballah, S.A.; Rizvi, T.A. Cross-packaging of genetically distinct mouse and primate retroviral RNAs. Retrovirology 2009, 6, 66. [Google Scholar] [CrossRef] [PubMed]
  26. Motomura, K.; Chen, J.; Hu, W.-S. Genetic Recombination between Human Immunodeficiency Virus Type 1 (HIV-1) and HIV-2, Two Distinct Human Lentiviruses. J. Virol. 2008, 82, 1923–1933. [Google Scholar] [CrossRef] [PubMed]
  27. Moore, M.D.; Fu, W.; Nikolaitchik, O.; Chen, J.; Ptak, R.G.; Hu, W.-S. Dimer Initiation Signal of Human Immunodeficiency Virus Type 1: Its Role in Partner Selection during RNA Copackaging and Its Effects on Recombination. J. Virol. 2007, 81, 4002–4011. [Google Scholar] [CrossRef] [PubMed]
  28. Parveen, Z.; Mukhtar, M.; Goodrich, A.; Acheampong, E.; Dornburg, R.; Pomerantz, R.J. Cross-Packaging of Human Immunodeficiency Virus Type 1 Vector RNA by Spleen Necrosis Virus Proteins: Construction of a New Generation of Spleen Necrosis Virus-Derived Retroviral Vectors. J. Virol. 2004, 78, 6480–6488. [Google Scholar] [CrossRef] [PubMed]
  29. Rizvi, T.A.; Panganiban, A.T. Simian immunodeficiency virus RNA is efficiently encapsidated by human immunodeficiency virus type 1 particles. J. Virol. 1993, 67, 2681–2688. [Google Scholar] [PubMed]
  30. White, S.M.; Renda, M.; Nam, N.-Y.; Klimatcheva, E.; Zhu, Y.; Fisk, J.; Halterman, M.; Rimel, B.J.; Federoff, H.; Pandya, S.; et al. Lentivirus Vectors Using Human and Simian Immunodeficiency Virus Elements. J. Virol. 1999, 73, 2832–2840. [Google Scholar] [PubMed]
  31. Yin, P.D.; Hu, W.S. RNAs from genetically distinct retroviruses can copackage and exchange genetic information in vivo. J. Virol. 1997, 71, 6237–6242. [Google Scholar] [PubMed]
  32. Jewell, N.A.; Mansky, L.M. In the beginning: genome recognition, RNA encapsidation and the initiation of complex retrovirus assembly. J. Gen. Virol. 2000, 81, 1889–1899. [Google Scholar] [CrossRef] [PubMed]
  33. Jouvenet, N.; Lainé, S.; Pessel-Vivares, L.; Mougel, M. Cell biology of retroviral RNA packaging. RNA Biol. 2011, 8, 572–580. [Google Scholar] [CrossRef] [PubMed]
  34. Erlwein, O.; Bieniasz, P.D.; McClure, M.O. Sequences in pol are required for transfer of human foamy virus-based vectors. J. Virol. 1998, 72, 5510–5516. [Google Scholar] [PubMed]
  35. Heinkelein, M.; Schmidt, M.; Fischer, N.; Moebes, A.; Lindemann, D.; Enssle, J.; Rethwilm, A. Characterization of a cis-acting sequence in the Pol region required to transfer human foamy virus vectors. J. Virol. 1998, 72, 6307–6314. [Google Scholar] [PubMed]
  36. Wu, M.; Chari, S.; Yanchis, T.; Mergia, A. cis-Acting sequences required for simian foamy virus type 1 vectors. J. Virol. 1998, 72, 3451–3454. [Google Scholar] [PubMed]
  37. Sorge, J.; Ricci, W.; Hughes, S.H. cis-Acting RNA packaging locus in the 115-nucleotide direct repeat of Rous sarcoma virus. J. Virol. 1983, 48, 667–675. [Google Scholar] [PubMed]
  38. Aschoff, J.M.; Foster, D.; Coffin, J.M. Point mutations in the avian sarcoma/leukosis virus 3’ untranslated region result in a packaging defect. J. Virol. 1999, 73, 7421–7429. [Google Scholar] [PubMed]
  39. Yu, S.S.; Kim, J.M.; Kim, S. The 17 nucleotides downstream from the env gene stop codon are important for murine leukemia virus packaging. J. Virol. 2000, 74, 8775–8780. [Google Scholar] [PubMed]
  40. Kaye, J.F.; Lever, A.M.L. Nonreciprocal Packaging of Human Immunodeficiency Virus Type 1 and Type 2 RNA: a Possible Role for the p2 Domain of Gag in RNA Encapsidation. J. Virol. 1998, 72, 5877–5885. [Google Scholar] [PubMed]
  41. Damgaard, C.K.; Dyhr-Mikkelsen, H.; Kjems, J. Mapping the RNA binding sites for human immunodeficiency virus type-1 gag and NC proteins within the complete HIV-1 and -2 untranslated leader regions. Nucleic Acids Res. 1998, 26, 3667–3676. [Google Scholar] [CrossRef] [PubMed]
  42. Griffin, S.D.; Allen, J.F.; Lever, A.M. The major human immunodeficiency virus type 2 (HIV-2) packaging signal is present on all HIV-2 RNA species: cotranslational RNA encapsidation and limitation of Gag protein confer specificity. J. Virol. 2001, 75, 12058–12069. [Google Scholar] [CrossRef] [PubMed]
  43. Baig, T.T.; Lanchy, J.-M.; Lodmell, J.S. Randomization and in vivo selection reveal a GGRG motif essential for packaging human immunodeficiency virus type 2 RNA. J. Virol. 2009, 83, 802–810. [Google Scholar] [CrossRef] [PubMed]
  44. Patel, J.; Wang, S.-W.; Izmailova, E.; Aldovini, A. The simian immunodeficiency virus 5’ untranslated leader sequence plays a role in intracellular viral protein accumulation and in RNA packaging. J. Virol. 2003, 77, 6284–6292. [Google Scholar] [CrossRef] [PubMed]
  45. Whitney, J.B.; Wainberg, M.A. Impaired RNA incorporation and dimerization in live attenuated leader-variants of SIVmac239. Retrovirology 2006, 3, 96. [Google Scholar] [CrossRef] [PubMed]
  46. Mustafa, F.; Ghazawi, A.; Jayanth, P.; Phillip, P.S.; Ali, J.; Rizvi, T.A. Sequences intervening between the core packaging determinants are dispensable for maintaining the packaging potential and propagation of feline immunodeficiency virus transfer vector RNAs. J. Virol. 2005, 79, 13817–13821. [Google Scholar] [CrossRef] [PubMed]
  47. Jaballah, S.A.; Aktar, S.J.; Ali, J.; Phillip, P.S.; Al Dhaheri, N.S.; Jabeen, A.; Rizvi, T.A. A G–C-Rich Palindromic Structural Motif and a Stretch of Single-Stranded Purines Are Required for Optimal Packaging of Mason–Pfizer Monkey Virus (MPMV) Genomic RNA. J. Mol. Biol. 2010, 401, 996–1014. [Google Scholar] [CrossRef] [PubMed]
  48. Schmidt, R.D.; Mustafa, F.; Lew, K.A.; Browning, M.T.; Rizvi, T.A. Sequences within both the 5’ untranslated region and the gag gene are important for efficient encapsidation of Mason-Pfizer monkey virus RNA. Virology 2003, 309, 166–178. [Google Scholar] [CrossRef]
  49. Mustafa, F.; Al Amri, D.; Al Ali, F.; Al Sari, N.; Al Suwaidi, S.; Jayanth, P.; Philips, P.S.; Rizvi, T.A. Sequences within both the 5’ UTR and Gag are required for optimal in vivo packaging and propagation of mouse mammary tumor virus (MMTV) genomic RNA. PLoS ONE 2012, 7, e47088. [Google Scholar] [CrossRef] [PubMed]
  50. Mansky, L.M.; Krueger, A.E.; Temin, H.M. The bovine leukemia virus encapsidation signal is discontinuous and extends into the 5’ end of the gag gene. J. Virol. 1995, 69, 3282–3289. [Google Scholar] [PubMed]
  51. Jewell, N.A.; Mansky, L.M. Packaging of heterologous RNAs by a minimal bovine leukemia virus RNA packaging signal into virus particles. Arch. Virol. 2005, 150, 1161–1173. [Google Scholar] [CrossRef] [PubMed]
  52. Mansky, L.M.; Wisniewski, R.M. The bovine leukemia virus encapsidation signal is composed of RNA secondary structures. J. Virol. 1998, 72, 3196–3204. [Google Scholar]
  53. Mansky, L.M.; Gajary, L.C. The primary nucleotide sequence of the bovine leukemia virus RNA packaging signal can influence efficient RNA packaging and virus replication. Virology 2002, 301, 272–280. [Google Scholar] [CrossRef] [PubMed]
  54. Mann, R.; Mulligan, R.C.; Baltimore, D. Construction of a retrovirus packaging mutant and its use to produce helper-free defective retrovirus. Cell 1983, 33, 153–159. [Google Scholar] [CrossRef]
  55. Adam, M.A.; Miller, A.D. Identification of a signal in a murine retrovirus that is sufficient for packaging of nonretroviral RNA into virions. J. Virol. 1988, 62, 3802–3806. [Google Scholar] [PubMed]
  56. Miyazaki, Y.; Irobalieva, R.N.; Tolbert, B.S.; Smalls-Mantey, A.; Iyalla, K.; Loeliger, K.; D’Souza, V.; Khant, H.; Schmid, M.F.; Garcia, E.L.; et al. Structure of a conserved retroviral RNA packaging element by NMR spectroscopy and cryo-electron tomography. J. Mol. Biol. 2010, 404, 751–772. [Google Scholar] [CrossRef] [PubMed]
  57. Watanabe, S.; Temin, H.M. Encapsidation sequences for spleen necrosis virus, an avian retrovirus, are between the 5’ long terminal repeat and the start of the gag gene. Proc. Natl. Acad. Sci. USA 1982, 79, 5986–5990. [Google Scholar] [CrossRef] [PubMed]
  58. Yang, S.; Temin, H.M. A double hairpin structure is necessary for the efficient encapsidation of spleen necrosis virus retroviral RNA. EMBO J. 1994, 13, 713–726. [Google Scholar] [PubMed]
  59. Katz, R.A.; Cullen, B.R.; Malavarca, R.; Skalka, A.M. Role of the avian retrovirus mRNA leader in expression: evidence for novel translational control. Mol. Cell. Biol. 1986, 6, 372–379. [Google Scholar] [CrossRef] [PubMed]
  60. Banks, J.D.; Linial, M.L. Secondary Structure Analysis of a Minimal Avian Leukosis-Sarcoma Virus Packaging Signal. J. Virol. 2000, 74, 456–464. [Google Scholar] [CrossRef] [PubMed]
  61. Zhou, J.; Bean, R.L.; Vogt, V.M.; Summers, M.F. Solution Structure of the Rous Sarcoma Virus Nucleocapsid Protein:μΨ RNA Packaging Signal Complex. J. Mol. Biol. 2007, 365, 453–467. [Google Scholar] [CrossRef] [PubMed]
  62. Konings, D.A.; Nash, M.A.; Maizel, J.V.; Arlinghaus, R.B. Novel GACG-hairpin pair motif in the 5’ untranslated region of type C retroviruses related to murine leukemia virus. J. Virol. 1992, 66, 632–640. [Google Scholar] [PubMed]
  63. Lever, A.; Gottlinger, H.; Haseltine, W.; Sodroski, J. Identification of a sequence required for efficient packaging of human immunodeficiency virus type 1 RNA into virions. J. Virol. 1989, 63, 4085–4087. [Google Scholar] [PubMed]
  64. Clavel, F.; Orenstein, J.M. A mutant of human immunodeficiency virus with reduced RNA packaging and abnormal particle morphology. J. Virol. 1990, 64, 5230–5234. [Google Scholar] [PubMed]
  65. Aldovini, A.; Young, R.A. Mutations of RNA and protein sequences involved in human immunodeficiency virus type 1 packaging result in production of noninfectious virus. J. Virol. 1990, 64, 1920–1926. [Google Scholar] [PubMed]
  66. Parolin, C.; Dorfman, T.; Palú, G.; Göttlinger, H.; Sodroski, J. Analysis in human immunodeficiency virus type 1 vectors of cis-acting sequences that affect gene transfer into human lymphocytes. J. Virol. 1994, 68, 3888–3895. [Google Scholar] [PubMed]
  67. Bender, M.A.; Palmer, T.D.; Gelinas, R.E.; Miller, A.D. Evidence that the packaging signal of Moloney murine leukemia virus extends into the gag region. J. Virol. 1987, 61, 1639–1646. [Google Scholar] [PubMed]
  68. Das, A.T.; Klaver, B.; Klasens, B.I.; van Wamel, J.L.; Berkhout, B. A conserved hairpin motif in the R-U5 region of the human immunodeficiency virus type 1 RNA genome is essential for replication. J. Virol. 1997, 71, 2346–2356. [Google Scholar]
  69. Aktar, S.J.; Jabeen, A.; Ali, L.M.; Vivet-Boudou, V.; Marquet, R.; Rizvi, T.A. SHAPE analysis of the 5’ end of the Mason-Pfizer monkey virus (MPMV) genomic RNA reveals structural elements required for genome dimerization. RNA 2013, 19, 1648–1658. [Google Scholar] [CrossRef] [PubMed]
  70. Aktar, S.J.; Vivet-Boudou, V.; Ali, L.M.; Jabeen, A.; Kalloush, R.M.; Richer, D.; Mustafa, F.; Marquet, R.; Rizvi, T.A. Structural basis of genomic RNA (gRNA) dimerization and packaging determinants of mouse mammary tumor virus (MMTV). Retrovirology 2014, 11, 96. [Google Scholar] [CrossRef] [PubMed]
  71. Paillart, J.-C.; Skripkin, E.; Ehresmann, B.; Ehresmann, C.; Marquet, R. In vitro evidence for a long range pseudoknot in the 5’-untranslated and matrix coding regions of HIV-1 genomic RNA. J. Biol. Chem. 2002, 277, 5995–6004. [Google Scholar] [CrossRef] [PubMed]
  72. Abbink, T.E.M.; Berkhout, B. A novel long distance base-pairing interaction in human immunodeficiency virus type 1 RNA occludes the Gag start codon. J. Biol. Chem. 2003, 278, 11601–11611. [Google Scholar] [CrossRef] [PubMed]
  73. Kenyon, J.C.; Ghazawi, A.; Cheung, W.K.S.; Phillip, P.S.; Rizvi, T.A.; Lever, A.M.L. The secondary structure of the 5’ end of the FIV genome reveals a long-range interaction between R/U5 and gag sequences, and a large, stable stem-loop. RNA 2008, 14, 2597–2608. [Google Scholar] [CrossRef] [PubMed]
  74. Kenyon, J.C.; Tanner, S.J.; Legiewicz, M.; Phillip, P.S.; Rizvi, T.A.; Le Grice, S.F.J.; Lever, A.M.L. SHAPE analysis of the FIV Leader RNA reveals a structural switch potentially controlling viral packaging and genome dimerization. Nucleic Acids Res. 2011, 39, 6692–6704. [Google Scholar] [CrossRef] [PubMed]
  75. Siegfried, N.A.; Busan, S.; Rice, G.M.; Nelson, J.A.E.; Weeks, K.M. RNA motif discovery by SHAPE and mutational profiling (SHAPE-MaP). Nat. Methods 2014, 11, 959–965. [Google Scholar] [CrossRef] [PubMed]
  76. Smyth, R.P.; Despons, L.; Huili, G.; Bernacchi, S.; Hijnen, M.; Mak, J.; Jossinet, F.; Weixi, L.; Paillart, J.-C.; von Kleist, M.; Marquet, R. Mutational interference mapping experiment (MIME) for studying RNA structure and function. Nat. Methods 2015, 12, 866–872. [Google Scholar] [CrossRef] [PubMed]
  77. Keane, S.C.; Heng, X.; Lu, K.; Kharytonchyk, S.; Ramakrishnan, V.; Carter, G.; Barton, S.; Hosic, A.; Florwick, A.; Santos, J.; et al. RNA structure. Structure of the HIV-1 RNA packaging signal. Science 2015, 348, 917–921. [Google Scholar] [CrossRef] [PubMed]
  78. Tran, T.; Liu, Y.; Marchant, J.; Monti, S.; Seu, M.; Zaki, J.; Yang, A.L.; Bohn, J.; Ramakrishnan, V.; Singh, R.; et al. Conserved determinants of lentiviral genome dimerization. Retrovirology 2015, 12, 83. [Google Scholar] [CrossRef] [PubMed]
  79. Kalloush, R.M.; Vivet-Boudou, V.; Ali, L.M.; Mustafa, F.; Marquet, R.; Rizvi, T.A. Packaging of Mason-Pfizer monkey virus (MPMV) genomic RNA depends upon conserved long-range interactions (LRIs) between U5 and gag sequences. RNA 2016, 22, 905–919. [Google Scholar] [CrossRef] [PubMed]
  80. Archer, E.J.; Simpson, M.A.; Watts, N.J.; O’Kane, R.; Wang, B.; Erie, D.A.; McPherson, A.; Weeks, K.M. Long-Range Architecture in a Viral RNA Genome. Biochemistry 2013, 52, 3182–3190. [Google Scholar] [CrossRef] [PubMed]
  81. Song, R.; Kafaie, J.; Laughrea, M. Role of the 5’ TAR stem--loop and the U5-AUG duplex in dimerization of HIV-1 genomic RNA. Biochemistry 2008, 47, 3283–3293. [Google Scholar] [CrossRef] [PubMed]
  82. Rizvi, T.A.; Kenyon, J.C.; Ali, J.; Aktar, S.J.; Phillip, P.S.; Ghazawi, A.; Mustafa, F.; Lever, A.M.L. Optimal Packaging of FIV Genomic RNA Depends upon a Conserved Long-range Interaction and a Palindromic Sequence within gag. J. Mol. Biol. 2010, 403, 103–119. [Google Scholar] [CrossRef] [PubMed]
  83. Abd El-Wahab, E.W.; Smyth, R.P.; Mailler, E.; Bernacchi, S.; Vivet-Boudou, V.; Hijnen, M.; Jossinet, F.; Mak, J.; Paillart, J.-C.; Marquet, R. Specific recognition of the HIV-1 genomic RNA by the Gag precursor. Nat. Commun. 2014, 5, 4304. [Google Scholar] [CrossRef] [PubMed]
  84. Linial, M.L.; Miller, A.D. Retroviral RNA packaging: sequence requirements and implications. Curr. Top. Microbiol. Immunol. 1990, 157, 125–152. [Google Scholar] [PubMed]
  85. Guesdon, F.M.J.; Greatorex, J.; Rhee, S.R.; Fisher, R.; Hunter, E.; Lever, A.M.L. Sequences in the 5′ Leader of Mason-Pfizer Monkey Virus Which Affect Viral Particle Production and Genomic RNA Packaging: Development of MPMV Packaging Cell Lines. Virology 2001, 288, 81–88. [Google Scholar] [CrossRef] [PubMed]
  86. Strappe, P.M.; Greatorex, J.; Thomas, J.; Biswas, P.; McCann, E.; Lever, A.M.L. The packaging signal of simian immunodeficiency virus is upstream of the major splice donor at a distance from the RNA cap site similar to that of human immunodeficiency virus types 1 and 2. J. Gen. Virol. 2003, 84, 2423–2430. [Google Scholar] [CrossRef] [PubMed]
  87. Balvay, L.; Lastra, M.L.; Sargueil, B.; Darlix, J.-L.; Ohlmann, T. Translational control of retroviruses. Nat. Rev. Microbiol. 2007, 5, 128–140. [Google Scholar] [CrossRef] [PubMed]
  88. Linial, M.; Brown, S.; Neiman, P. A nonconditional mutant of Rous sarcoma virus containing defective polymerase. Virology 1978, 87, 130–141. [Google Scholar] [CrossRef]
  89. Donzé, O.; Spahr, P.F. Role of the open reading frames of Rous sarcoma virus leader RNA in translation and genome packaging. EMBO J. 1992, 11, 3747–3757. [Google Scholar] [PubMed]
  90. Donzé, O.; Damay, P.; Spahr, P.F. The first and third uORFs in RSV leader RNA are efficiently translated: implications for translational regulation and viral RNA packaging. Nucleic Acids Res. 1995, 23, 861–868. [Google Scholar] [CrossRef] [PubMed]
  91. Kaye, J.F.; Lever, A.M. Human immunodeficiency virus types 1 and 2 differ in the predominant mechanism used for selection of genomic RNA for encapsidation. J. Virol. 1999, 73, 3023–3031. [Google Scholar] [PubMed]
  92. Luban, J.; Goff, S.P. Mutational analysis of cis-acting packaging signals in human immunodeficiency virus type 1 RNA. J. Virol. 1994, 68, 3784–3793. [Google Scholar] [PubMed]
  93. Ott, D.E.; Coren, L.V.; Sowder, R.C.; Adams, J.; Schubert, U. Retroviruses have differing requirements for proteasome function in the budding process. J. Virol. 2003, 77, 3384–3393. [Google Scholar] [CrossRef] [PubMed]
  94. Ott, D.E.; Coren, L.V.; Gagliardi, T.D.; Nagashima, K. Heterologous late-domain sequences have various abilities to promote budding of human immunodeficiency virus type 1. J. Virol. 2005, 79, 9038–9045. [Google Scholar] [CrossRef] [PubMed]
  95. Swanstrom, R.; Wills, J.W. Synthesis, Assembly, and Processing of Viral Proteins. In Retroviruses; Coffin, J.M., Hughes, S.H., Varmus, H.E., Eds.; Cold Spring Harbor Laboratory Press: Cold Spring Harbor, NY, USA, 1997. [Google Scholar]
  96. Vogt, V.M. Retroviral Virions and Genomes. In Retroviruses; Coffin, J.M., Hughes, S.H., Varmus, H.E., Eds.; Cold Spring Harbor Laboratory Press: Cold Spring Harbor, NY, USA, 1997. [Google Scholar]
  97. Green, L.M.; Berg, J.M. A retroviral Cys-Xaa2-Cys-Xaa4-His-Xaa4-Cys peptide binds metal ions: spectroscopic studies and a proposed three-dimensional structure. Proc. Natl. Acad. Sci. USA 1989, 86, 4047–4051. [Google Scholar] [CrossRef] [PubMed]
  98. Henderson, L.E.; Copeland, T.D.; Sowder, R.C.; Smythers, G.W.; Oroszlan, S. Primary structure of the low molecular weight nucleic acid-binding proteins of murine leukemia viruses. J. Biol. Chem. 1981, 256, 8400–8406. [Google Scholar] [PubMed]
  99. Bess, J.W.; Powell, P.J.; Issaq, H.J.; Schumack, L.J.; Grimes, M.K.; Henderson, L.E.; Arthur, L.O. Tightly bound zinc in human immunodeficiency virus type 1, human T-cell leukemia virus type I, and other retroviruses. J. Virol. 1992, 66, 840–847. [Google Scholar] [PubMed]
  100. Chance, M.R.; Sagi, I.; Wirt, M.D.; Frisbie, S.M.; Scheuring, E.; Chen, E.; Bess, J.W.; Henderson, L.E.; Arthur, L.O.; South, T.L. Extended x-ray absorption fine structure studies of a retrovirus: equine infectious anemia virus cysteine arrays are coordinated to zinc. Proc. Natl. Acad. Sci. USA 1992, 89, 10041–10045. [Google Scholar] [CrossRef] [PubMed]
  101. Summers, M.F.; Henderson, L.E.; Chance, M.R.; Bess, J.W.; South, T.L.; Blake, P.R.; Sagi, I.; Perez-Alvarado, G.; Sowder, R.C.; Hare, D.R. Nucleocapsid zinc fingers detected in retroviruses: EXAFS studies of intact viruses and the solution-state structure of the nucleocapsid protein from HIV-1. Protein Sci. 1992, 1, 563–574. [Google Scholar] [CrossRef] [PubMed]
  102. Poon, D.T.; Wu, J.; Aldovini, A. Charged amino acid residues of human immunodeficiency virus type 1 nucleocapsid p7 protein involved in RNA packaging and infectivity. J. Virol. 1996, 70, 6607–6616. [Google Scholar] [PubMed]
  103. Gorelick, R.J.; Henderson, L.E.; Hanser, J.P.; Rein, A. Point mutants of Moloney murine leukemia virus that fail to package viral RNA: evidence for specific RNA recognition by a “zinc finger-like” protein sequence. Proc. Natl. Acad. Sci. USA 1988, 85, 8420–8424. [Google Scholar] [CrossRef] [PubMed]
  104. Jentoft, J.E.; Smith, L.M.; Fu, X.D.; Johnson, M.; Leis, J. Conserved cysteine and histidine residues of the avian myeloblastosis virus nucleocapsid protein are essential for viral replication but are not “zinc-binding fingers.”. Proc. Natl. Acad. Sci. USA 1988, 85, 7094–7098. [Google Scholar] [CrossRef] [PubMed]
  105. Méric, C.; Gouilloud, E.; Spahr, P.F. Mutations in Rous sarcoma virus nucleocapsid protein p12 (NC): deletions of Cys-His boxes. J. Virol. 1988, 62, 3328–3333. [Google Scholar] [PubMed]
  106. Gorelick, R.J.; Nigida, S.M.J.; Bess, J.W.; Arthur, L.O.; Henderson, L.E.; Rein, A. Noninfectious human immunodeficiency virus type 1 mutants deficient in genomic RNA. J. Virol. 1990, 64, 3207–11. [Google Scholar] [PubMed]
  107. Dorfman, T.; Luban, J.; Goff, S.P.; Haseltine, W.A.; Göttlinger, H.G. Mapping of functionally important residues of a cysteine-histidine box in the human immunodeficiency virus type 1 nucleocapsid protein. J. Virol. 1993, 67, 6159–6169. [Google Scholar] [PubMed]
  108. Gorelick, R.J.; Chabot, D.J.; Rein, A.; Henderson, L.E.; Arthur, L.O. The two zinc fingers in the human immunodeficiency virus type 1 nucleocapsid protein are not functionally equivalent. J. Virol. 1993, 67, 4027–4036. [Google Scholar] [PubMed]
  109. Wang, C.T.; Barklis, E. Assembly, processing, and infectivity of human immunodeficiency virus type 1 Gag mutants. J. Virol. 1993, 67, 4264–4273. [Google Scholar] [PubMed]
  110. Zhang, Y.; Barklis, E. Nucleocapsid protein effects on the specificity of retrovirus RNA encapsidation. J. Virol. 1995, 69, 5716–5722. [Google Scholar] [PubMed]
  111. Mizuno, A.; Ido, E.; Goto, T.; Kuwata, T.; Nakai, M.; Hayami, M. Mutational analysis of two zinc finger motifs in HIV type 1 nucleocapsid proteins: effects on proteolytic processing of Gag precursors and particle formation. AIDS Res. Hum. Retroviruses 1996, 12, 793–800. [Google Scholar] [CrossRef] [PubMed]
  112. De Guzman, R.N.; Turner, R.B.; Summers, M.F. Protein-RNA recognition. Biopolymers 1998, 48, 181–195. [Google Scholar] [CrossRef]
  113. Tanchou, V.; Decimo, D.; Péchoux, C.; Lener, D.; Rogemond, V.; Berthoux, L.; Ottmann, M.; Darlix, J.L. Role of the N-terminal zinc finger of human immunodeficiency virus type 1 nucleocapsid protein in virus structure and replication. J. Virol. 1998, 72, 4442–4447. [Google Scholar] [PubMed]
  114. Gorelick, R.J.; Gagliardi, T.D.; Bosche, W.J.; Wiltrout, T.A.; Coren, L.V.; Chabot, D.J.; Lifson, J.D.; Henderson, L.E.; Arthur, L.O. Strict conservation of the retroviral nucleocapsid protein zinc finger is strongly influenced by its role in viral infection processes: characterization of HIV-1 particles containing mutant nucleocapsid zinc-coordinating sequences. Virology 1999, 256, 92–104. [Google Scholar] [CrossRef] [PubMed]
  115. Guo, J.; Wu, T.; Anderson, J.; Kane, B.F.; Johnson, D.G.; Gorelick, R.J.; Henderson, L.E.; Levin, J.G. Zinc finger structures in the human immunodeficiency virus type 1 nucleocapsid protein facilitate efficient minus- and plus-strand transfer. J. Virol. 2000, 74, 8980–8988. [Google Scholar] [CrossRef] [PubMed]
  116. Housset, V.; De Rocquigny, H.; Roques, B.P.; Darlix, J.L. Basic amino acids flanking the zinc finger of Moloney murine leukemia virus nucleocapsid protein NCp10 are critical for virus infectivity. J. Virol. 1993, 67, 2537–2545. [Google Scholar] [PubMed]
  117. Cimarelli, A.; Sandin, S.; Höglund, S.; Luban, J. Basic residues in human immunodeficiency virus type 1 nucleocapsid promote virion assembly via interaction with RNA. J. Virol. 2000, 74, 3046–3057. [Google Scholar] [CrossRef] [PubMed]
  118. Gonsky, J.; Bacharach, E.; Goff, S.P. Identification of residues of the Moloney murine leukemia virus nucleocapsid critical for viral DNA synthesis in vivo. J. Virol. 2001, 75, 2616–2626. [Google Scholar] [CrossRef] [PubMed]
  119. Dupraz, P.; Spahr, P.F. Specificity of Rous sarcoma virus nucleocapsid protein in genomic RNA packaging. J. Virol. 1992, 66, 4662–4670. [Google Scholar] [PubMed]
  120. Berkowitz, R.D.; Ohagen, A.; Höglund, S.; Goff, S.P. Retroviral nucleocapsid domains mediate the specific recognition of genomic viral RNAs by chimeric Gag polyproteins during RNA packaging in vivo. J. Virol. 1995, 69, 6445–6456. [Google Scholar] [PubMed]
  121. Fu, W.; Hu, W.-S. Functional Replacement of Nucleocapsid Flanking Regions by Heterologous Counterparts with Divergent Primary Sequences: Effects of Chimeric Nucleocapsid on the Retroviral Replication Cycle. J. Virol. 2003, 77, 754–761. [Google Scholar] [CrossRef] [PubMed]
  122. Freed, E.O. HIV-1 gag proteins: diverse functions in the virus life cycle. Virology 1998, 251, 1–15. [Google Scholar] [CrossRef] [PubMed]
  123. Miyazaki, Y.; Garcia, E.L.; King, S.R.; Iyalla, K.; Loeliger, K.; Starck, P.; Syed, S.; Telesnitsky, A.; Summers, M.F. An RNA structural switch regulates diploid genome packaging by Moloney murine leukemia virus. J. Mol. Biol. 2010, 396, 141–152. [Google Scholar] [CrossRef] [PubMed]
  124. D’Souza, V.; Summers, M.F. Structural basis for packaging the dimeric genome of Moloney murine leukaemia virus. Nature 2004, 431, 586–590. [Google Scholar] [CrossRef] [PubMed]
  125. Poon, D.T.; Li, G.; Aldovini, A. Nucleocapsid and matrix protein contributions to selective human immunodeficiency virus type 1 genomic RNA packaging. J. Virol. 1998, 72, 1983–1993. [Google Scholar] [PubMed]
  126. Guo, X.; Roy, B.B.; Hu, J.; Roldan, A.; Wainberg, M.A.; Liang, C. The R362A mutation at the C-terminus of CA inhibits packaging of human immunodeficiency virus type 1 RNA. Virology 2005, 343, 190–200. [Google Scholar] [CrossRef] [PubMed]
  127. Russell, R.S.; Roldan, A.; Detorio, M.; Hu, J.; Wainberg, M.A.; Liang, C. Effects of a single amino acid substitution within the p2 region of human immunodeficiency virus type 1 on packaging of spliced viral RNA. J. Virol. 2003, 77, 12986–12995. [Google Scholar] [CrossRef] [PubMed]
  128. Roy, B.B.; Russell, R.S.; Turner, D.; Liang, C. The T12I mutation within the SP1 region of Gag restricts packaging of spliced viral RNA into human immunodeficiency virus type 1 with mutated RNA packaging signals and mutated nucleocapsid sequence. Virology 2006, 344, 304–314. [Google Scholar] [CrossRef] [PubMed]
  129. Butterfield-Gerson, K.L.; Scheifele, L.Z.; Ryan, E.P.; Hopper, A.K.; Parent, L.J. Importin-beta family members mediate alpharetrovirus gag nuclear entry via interactions with matrix and nucleocapsid. J. Virol. 2006, 80, 1798–1806. [Google Scholar] [CrossRef] [PubMed]
  130. Kenney, S.P.; Lochmann, T.L.; Schmid, C.L.; Parent, L.J. Intermolecular interactions between retroviral Gag proteins in the nucleus. J. Virol. 2008, 82, 683–691. [Google Scholar] [CrossRef] [PubMed]
  131. Gudleski, N.; Flanagan, J.M.; Ryan, E.P.; Bewley, M.C.; Parent, L.J. Directionality of nucleocytoplasmic transport of the retroviral gag protein depends on sequential binding of karyopherins and viral RNA. Proc. Natl. Acad. Sci. USA 2010, 107, 9358–9363. [Google Scholar] [CrossRef] [PubMed]
  132. Cullen, B.R. Nuclear mRNA export: insights from virology. Trends Biochem. Sci. 2003, 28, 419–424. [Google Scholar] [CrossRef]
  133. Jouvenet, N.; Simon, S.M.; Bieniasz, P.D. Imaging the interaction of HIV-1 genomes and Gag during assembly of individual viral particles. Proc. Natl. Acad. Sci. USA 2009, 106, 19114–19119. [Google Scholar] [CrossRef] [PubMed]
  134. Kemler, I.; Meehan, A.; Poeschla, E.M. Live-cell coimaging of the genomic RNAs and Gag proteins of two lentiviruses. J. Virol. 2010, 84, 6352–6366. [Google Scholar] [CrossRef] [PubMed]
  135. Saad, J.S.; Miller, J.; Tai, J.; Kim, A.; Ghanam, R.H.; Summers, M.F. Structural basis for targeting HIV-1 Gag proteins to the plasma membrane for virus assembly. Proc. Natl. Acad. Sci. USA 2006, 103, 11364–11369. [Google Scholar] [CrossRef] [PubMed]
  136. Ono, A.; Ablan, S.D.; Lockett, S.J.; Nagashima, K.; Freed, E.O. Phosphatidylinositol (4,5) bisphosphate regulates HIV-1 Gag targeting to the plasma membrane. Proc. Natl. Acad. Sci. USA 2004, 101, 14889–14894. [Google Scholar] [CrossRef] [PubMed]
  137. Strappe, P.M.; Hampton, D.W.; Brown, D.; Cachon-Gonzalez, B.; Caldwell, M.; Fawcett, J.W.; Lever, A.M. Identification of unique reciprocal and non-reciprocal cross packaging relationships between HIV-1, HIV-2 and SIV reveals an efficient SIV/HIV-2 lentiviral vector system with highly favourable features for in vivo testing and clinical usage. Retrovirology 2005, 2, 55. [Google Scholar] [CrossRef] [PubMed]
  138. Certo, J.L.; Shook, B.F.; Yin, P.D.; Snider, J.T.; Hu, W.-S. Nonreciprocal Pseudotyping: Murine Leukemia Virus Proteins Cannot Efficiently Package Spleen Necrosis Virus-Based Vector RNA. J. Virol. 1998, 72, 5408–5413. [Google Scholar] [PubMed]
  139. Kewalramani, V.N.; Panganiban, A.T.; Emerman, M. Spleen necrosis virus, an avian immunosuppressive retrovirus, shares a receptor with the type D simian retroviruses. J. Virol. 1992, 66, 3026–3031. [Google Scholar] [PubMed]
  140. Dougherty, J.P.; Wisniewski, R.; Yang, S.L.; Rhode, B.W.; Temin, H.M. New retrovirus helper cells with almost no nucleotide sequence homology to retrovirus vectors. J. Virol. 1989, 63, 3209–3212. [Google Scholar] [PubMed]
  141. Embretson, J.E.; Temin, H.M. Lack of competition results in efficient packaging of heterologous murine retroviral RNAs and reticuloendotheliosis virus encapsidation-minus RNAs by the reticuloendotheliosis virus helper cell line. J. Virol. 1987, 61, 2675–2683. [Google Scholar] [PubMed]
  142. Chakrabarti, R.; Hofman, F.M.; Pandey, R.; Mathes, L.E.; Roy-Burman, P. Recombination between feline exogenous and endogenous retroviral sequences generates tropism for cerebral endothelial cells. Am. J. Pathol. 1994, 144, 348–358. [Google Scholar] [PubMed]
  143. Chong, H.; Starkey, W.; Vile, R.G. A replication-competent retrovirus arising from a split-function packaging cell line was generated by recombination events between the vector, one of the packaging constructs, and endogenous retroviral sequences. J. Virol. 1998, 72, 2663–2670. [Google Scholar] [PubMed]
  144. Bukovsky, A.A.; Song, J.P.; Naldini, L. Interaction of human immunodeficiency virus-derived vectors with wild-type virus in transduced cells. J. Virol. 1999, 73, 7087–7092. [Google Scholar] [PubMed]
  145. Bartosch, B.; Stefanidis, D.; Myers, R.; Weiss, R.; Patience, C.; Takeuchi, Y. Evidence and consequence of porcine endogenous retrovirus recombination. J. Virol. 2004, 78, 13880–13890. [Google Scholar] [CrossRef] [PubMed]
  146. Evans, J.T.; Garcia, J.V. Lentivirus vector mobilization and spread by human immunodeficiency virus. Hum. Gene Ther. 2000, 11, 2331–2339. [Google Scholar] [CrossRef] [PubMed]
  147. Evans, L.H.; Alamgir, A.S.M.; Owens, N.; Weber, N.; Virtaneva, K.; Barbian, K.; Babar, A.; Malik, F.; Rosenke, K. Mobilization of endogenous retroviruses in mice after infection with an exogenous retrovirus. J. Virol. 2009, 83, 2429–2435. [Google Scholar] [CrossRef] [PubMed]
  148. Dull, T.; Zufferey, R.; Kelly, M.; Mandel, R.J.; Nguyen, M.; Trono, D.; Naldini, L. A third-generation lentivirus vector with a conditional packaging system. J. Virol. 1998, 72, 8463–8471. [Google Scholar] [PubMed]
  149. Ismail, S.I.; Kingsman, S.M.; Kingsman, A.J.; Uden, M. Split-intron retroviral vectors: enhanced expression with improved safety. J. Virol. 2000, 74, 2365–2371. [Google Scholar] [CrossRef] [PubMed]
  150. Zufferey, R.; Dull, T.; Mandel, R.J.; Bukovsky, A.; Quiroz, D.; Naldini, L.; Trono, D. Self-inactivating lentivirus vector for safe and efficient in vivo gene delivery. J. Virol. 1998, 72, 9873–9880. [Google Scholar] [PubMed]
  151. Westerman, K.A.; Ao, Z.; Cohen, E.A.; Leboulch, P. Design of a trans protease lentiviral packaging system that produces high titer virus. Retrovirology 2007, 4, 96. [Google Scholar] [CrossRef] [PubMed]
  152. Logan, A.C.; Haas, D.L.; Kafri, T.; Kohn, D.B. Integrated self-inactivating lentiviral vectors produce full-length genomic transcripts competent for encapsidation and integration. J. Virol. 2004, 78, 8421–8436. [Google Scholar] [CrossRef] [PubMed]
  153. Hanawa, H.; Persons, D.A.; Nienhuis, A.W. Mobilization and mechanism of transcription of integrated self-inactivating lentiviral vectors. J. Virol. 2005, 79, 8410–8421. [Google Scholar] [CrossRef] [PubMed]
  154. Schambach, A.; Galla, M.; Maetzig, T.; Loew, R.; Baum, C. Improving transcriptional termination of self-inactivating gamma-retroviral and lentiviral vectors. Mol. Ther. J. Am. Soc. Gene Ther. 2007, 15, 1167–1173. [Google Scholar] [CrossRef] [PubMed]
  155. Goujon, C.; Jarrosson-Wuilleme, L.; Bernaud, J.; Rigal, D.; Darlix, J.-L.; Cimarelli, A. Heterologous human immunodeficiency virus type 1 lentiviral vectors packaging a simian immunodeficiency virus-derived genome display a specific postentry transduction defect in dendritic cells. J. Virol. 2003, 77, 9295–9304. [Google Scholar] [CrossRef] [PubMed]
  156. Poeschla, E.M.; Wong-Staal, F.; Looney, D.J. Efficient transduction of nondividing human cells by feline immunodeficiency virus lentiviral vectors. Nat. Med. 1998, 4, 354–357. [Google Scholar] [CrossRef] [PubMed]
  157. Morris, K.V.; Gilbert, J.; Wong-Staal, F.; Gasmi, M.; Looney, D.J. Transduction of cell lines and primary cells by FIV-packaged HIV vectors. Mol. Ther. J. Am. Soc. Gene Ther. 2004, 10, 181–190. [Google Scholar] [CrossRef]
  158. Delviks, K.A.; Hu, W.S.; Pathak, V.K. Psi- vectors: murine leukemia virus-based self-inactivating and self-activating retroviral vectors. J. Virol. 1997, 71, 6218–6224. [Google Scholar] [PubMed]
  159. Saller, R.M.; Oztürk, F.; Salmons, B.; Günzburg, W.H. Construction and characterization of a hybrid mouse mammary tumor virus/murine leukemia virus-based retroviral vector. J. Virol. 1998, 72, 1699–1703. [Google Scholar] [PubMed]
  160. Beasley, B.E.; Hu, W.-S. cis-Acting elements important for retroviral RNA packaging specificity. J. Virol. 2002, 76, 4950–4960. [Google Scholar] [CrossRef] [PubMed]
  161. Browning, M.T.; Schmidt, R.D.; Lew, K.A.; Rizvi, T.A. Primate and feline lentivirus vector RNA packaging and propagation by heterologous lentivirus virions. J. Virol. 2001, 75, 5129–5140. [Google Scholar] [CrossRef] [PubMed]
  162. Kol, N.; Gladnikoff, M.; Barlam, D.; Shneck, R.Z.; Rein, A.; Rousso, I. Mechanical properties of murine leukemia virus particles: effect of maturation. Biophys. J. 2006, 91, 767–774. [Google Scholar] [CrossRef] [PubMed]
  163. Günzburg, W.H.; Salmons, B. Mouse mammary tumor virus mediated transfer and expression of neomycin resistance to infected cultured cells. Virology 1986, 155, 236–248. [Google Scholar] [CrossRef]
  164. Masuda, T.; Kuroda, M.J.; Harada, S. Specific and independent recognition of U3 and U5 att sites by human immunodeficiency virus type 1 integrase in vivo. J. Virol. 1998, 72, 8396–8402. [Google Scholar] [PubMed]
  165. Hlavaty, J.; Stracke, A.; Klein, D.; Salmons, B.; Günzburg, W.H.; Renner, M. Multiple modifications allow high-titer production of retroviral vectors carrying heterologous regulatory elements. J. Virol. 2004, 78, 1384–1392. [Google Scholar] [CrossRef] [PubMed]
  166. Ghazawi, A.; Mustafa, F.; Phillip, P.S.; Jayanth, P.; Ali, J.; Rizvi, T.A. Both the 5’ and 3’ LTRs of FIV contain minor RNA encapsidation determinants compared to the two core packaging determinants within the 5’ untranslated region and gag. Microbes Infect. 2006, 8, 767–778. [Google Scholar] [CrossRef] [PubMed]
  167. Adkins, B.; Hunter, T. Identification of a packaged cellular mRNA in virions of rous sarcoma virus. J. Virol. 1981, 39, 471–480. [Google Scholar] [PubMed]
  168. Aronoff, R.; Linial, M. Specificity of retroviral RNA packaging. J. Virol. 1991, 65, 71–80. [Google Scholar] [PubMed]
  169. Bishop, J.M.; Levinson, W.E.; Sullivan, D.; Fanshier, L.; Quintrell, N.; Jackson, J. The low molecular weight RNAs of Rous sarcoma virus. II. The 7 S RNA. Virology 1970, 42, 927–937. [Google Scholar] [CrossRef]
  170. Gallis, B.; Linial, M.; Eisenman, R. An avian oncovirus mutant deficient in genomic RNA: characterization of the packaged RNA as cellular messenger RNA. Virology 1979, 94, 146–161. [Google Scholar] [CrossRef]
  171. Ikawa, Y.; Ross, J.; Leder, P. An association between globin messenger RNA and 60S RNA derived from Friend leukemia virus. Proc. Natl. Acad. Sci. USA 1974, 71, 1154–1158. [Google Scholar] [CrossRef] [PubMed]
  172. Muriaux, D.; Mirro, J.; Harvin, D.; Rein, A. RNA is a structural element in retrovirus particles. Proc. Natl. Acad. Sci. USA 2001, 98, 5246–5251. [Google Scholar] [CrossRef] [PubMed]
  173. Waters, L.C.; Mullin, B.C. Transfer RNA into RNA tumor viruses. Prog. Nucleic Acid Res. Mol. Biol. 1977, 20, 131–160. [Google Scholar] [PubMed]
  174. Martin, P.; Henry, C.; Ferre, F.; Bechade, C.; Begue, A.; Calothy, C.; Debuire, B.; Stehelin, D.; Saule, S. Characterization of a myc-containing retrovirus generated by propagation of an MH2 viral subgenomic RNA. J. Virol. 1986, 57, 1191–1194. [Google Scholar] [PubMed]
  175. Stacey, D.W. Messenger activity of virion RNA for avian leukosis viral envelope glycoprotein. J. Virol. 1979, 29, 949–956. [Google Scholar] [PubMed]
  176. Svoboda, J.; Dvorák, M.; Guntaka, R.; Geryk, J. Transmission of (LTR, v-src, LTR) without recombination with a helper virus. Virology 1986, 153, 314–317. [Google Scholar] [CrossRef]
  177. Onafuwa-Nuga, A.A.; King, S.R.; Telesnitsky, A. Nonrandom packaging of host RNAs in moloney murine leukemia virus. J. Virol. 2005, 79, 13528–13537. [Google Scholar] [CrossRef] [PubMed]
  178. Eckwahl, M.J.; Telesnitsky, A.; Wolin, S.L. Host RNA Packaging by Retroviruses: A Newly Synthesized Story. MBio 2016, 7, e02025-15. [Google Scholar] [CrossRef] [PubMed]
  179. Onafuwa-Nuga, A.; Telesnitsky, A. The remarkable frequency of human immunodeficiency virus type 1 genetic recombination. Microbiol. Mol. Biol. Rev. 2009, 73, 451–480, Table of Contents. [Google Scholar] [CrossRef] [PubMed]
  180. Marquet, R.; Baudin, F.; Gabus, C.; Darlix, J.L.; Mougel, M.; Ehresmann, C.; Ehresmann, B. Dimerization of human immunodeficiency virus (type 1) RNA: stimulation by cations and possible mechanism. Nucleic Acids Res. 1991, 19, 2349–2357. [Google Scholar] [CrossRef] [PubMed]
  181. Delviks-Frankenberry, K.; Galli, A.; Nikolaitchik, O.; Mens, H.; Pathak, V.K.; Hu, W.-S. Mechanisms and Factors that Influence High Frequency Retroviral Recombination. Viruses 2011, 3, 1650–1680. [Google Scholar] [CrossRef] [PubMed]
  182. Negroni, M.; Buc, H. Mechanisms of retroviral recombination. Annu. Rev. Genet. 2001, 35, 275–302. [Google Scholar] [CrossRef] [PubMed]
  183. Ni, N.; Nikolaitchik, O.A.; Dilley, K.A.; Chen, J.; Galli, A.; Fu, W.; Prasad, V.V.S.P.; Ptak, R.G.; Pathak, V.K.; Hu, W.-S. Mechanisms of human immunodeficiency virus type 2 RNA packaging: efficient trans packaging and selection of RNA copackaging partners. J. Virol. 2011, 85, 7603–7612. [Google Scholar] [CrossRef] [PubMed]
  184. Dilley, K.A.; Ni, N.; Nikolaitchik, O.A.; Chen, J.; Galli, A.; Hu, W.-S. Determining the frequency and mechanisms of HIV-1 and HIV-2 RNA copackaging by single-virion analysis. J. Virol. 2011, 85, 10499–10508. [Google Scholar] [CrossRef] [PubMed]
  185. Mikkelsen, J.G.; Rasmussen, S.V.; Pedersen, F.S. Complementarity-directed RNA dimer-linkage promotes retroviral recombination in vivo. Nucleic Acids Res. 2004, 32, 102–114. [Google Scholar] [CrossRef] [PubMed]
  186. Rasmussen, S.V.; Mikkelsen, J.G.; Pedersen, F.S. Modulation of homo- and heterodimerization of Harvey sarcoma virus RNA by GACG tetraloops and point mutations in palindromic sequences. J. Mol. Biol. 2002, 323, 613–628. [Google Scholar] [CrossRef]
  187. Vogt, P.K. Genetically stable reassortment of markers during mixed infection with avian tumor viruses. Virology 1971, 46, 947–952. [Google Scholar] [CrossRef]
  188. Clavel, F.; Hoggan, M.D.; Willey, R.L.; Strebel, K.; Martin, M.A.; Repaske, R. Genetic recombination of human immunodeficiency virus. J. Virol. 1989, 63, 1455–1459. [Google Scholar] [PubMed]
  189. Wong, P.K.; Russ, L.J.; McCarter, J.A. Rapid, selective procedure for isolation of spontaneous temperature-sensitive mutants of Moloney leukemia virus. Virology 1973, 51, 424–431. [Google Scholar] [CrossRef]
  190. Wyke, J.A. Temperature sensitive mutants of avian sarcoma viruses. Biochim. Biophys. Acta 1975, 417, 91–121. [Google Scholar] [CrossRef]
  191. Dang, Q.; Chen, J.; Unutmaz, D.; Coffin, J.M.; Pathak, V.K.; Powell, D.; KewalRamani, V.N.; Maldarelli, F.; Hu, W.-S. Nonrandom HIV-1 infection and double infection via direct and cell-mediated pathways. Proc. Natl. Acad. Sci. USA 2004, 101, 632–637. [Google Scholar] [CrossRef] [PubMed]
  192. Chen, J.; Dang, Q.; Unutmaz, D.; Pathak, V.K.; Maldarelli, F.; Powell, D.; Hu, W.-S. Mechanisms of nonrandom human immunodeficiency virus type 1 infection and double infection: preference in virus entry is important but is not the sole factor. J. Virol. 2005, 79, 4140–4149. [Google Scholar] [CrossRef] [PubMed]
  193. Del Portillo, A.; Tripodi, J.; Najfeld, V.; Wodarz, D.; Levy, D.N.; Chen, B.K. Multiploid inheritance of HIV-1 during cell-to-cell infection. J. Virol. 2011, 85, 7169–7176. [Google Scholar] [CrossRef] [PubMed]
  194. Weiss, R.A.; Mason, W.S.; Vogt, P.K. Genetic recombinants and heterozygotes derived from endogenous and exogenous avian RNA tumor viruses. Virology 1973, 52, 535–552. [Google Scholar] [CrossRef]
  195. Hu, W.S.; Temin, H.M. Genetic consequences of packaging two RNA genomes in one retroviral particle: pseudodiploidy and high rate of genetic recombination. Proc. Natl. Acad. Sci. USA 1990, 87, 1556–1560. [Google Scholar] [CrossRef] [PubMed]
  196. Pathak, V.K.; Hu, W.-S. “Might as Well Jump!” Template Switching by Retroviral Reverse Transcriptase, Defective Genome Formation, and Recombination. Semin. Virol. 1997, 8, 141–150. [Google Scholar] [CrossRef]
  197. Gilboa, E.; Mitra, S.W.; Goff, S.; Baltimore, D. A detailed model of reverse transcription and tests of crucial aspects. Cell 1979, 18, 93–100. [Google Scholar] [CrossRef]
  198. Goff, S.P. Retroviral reverse transcriptase: synthesis, structure, and function. J. Acquir. Immune Defic. Syndr. 1990, 3, 817–831. [Google Scholar] [PubMed]
  199. Yu, H.; Jetzt, A.E.; Ron, Y.; Preston, B.D.; Dougherty, J.P. The nature of human immunodeficiency virus type 1 strand transfers. J. Biol. Chem. 1998, 273, 28384–28391. [Google Scholar] [CrossRef] [PubMed]
  200. van Wamel, J.L.; Berkhout, B. The first strand transfer during HIV-1 reverse transcription can occur either intramolecularly or intermolecularly. Virology 1998, 244, 245–251. [Google Scholar] [CrossRef] [PubMed]
  201. Stuhlmann, H.; Berg, P. Homologous recombination of copackaged retrovirus RNAs during reverse transcription. J. Virol. 1992, 66, 2378–2388. [Google Scholar] [PubMed]
  202. Hu, W.S.; Bowman, E.H.; Delviks, K.A.; Pathak, V.K. Homologous recombination occurs in a distinct retroviral subpopulation and exhibits high negative interference. J. Virol. 1997, 71, 6028–6036. [Google Scholar] [PubMed]
  203. Katz, R.A.; Skalka, A.M. Generation of diversity in retroviruses. Annu. Rev. Genet. 1990, 24, 409–445. [Google Scholar] [CrossRef] [PubMed]
  204. Anderson, J.A.; Teufel, R.J.; Yin, P.D.; Hu, W.S. Correlated template-switching events during minus-strand DNA synthesis: a mechanism for high negative interference during retroviral recombination. J. Virol. 1998, 72, 1186–1194. [Google Scholar] [PubMed]
  205. Zhang, J.; Tang, L.-Y.; Li, T.; Ma, Y.; Sapp, C.M. Most Retroviral Recombinations Occur during Minus-Strand DNA Synthesis. J. Virol. 2000, 74, 2313–2322. [Google Scholar] [CrossRef] [PubMed]
  206. Pathak, V.K.; Temin, H.M. Broad spectrum of in vivo forward mutations, hypermutations, and mutational hotspots in a retroviral shuttle vector after a single replication cycle: deletions and deletions with insertions. Proc. Natl. Acad. Sci. USA 1990, 87, 6024–6028. [Google Scholar] [CrossRef] [PubMed]
  207. Parthasarathi, S.; Varela-Echavarría, A.; Ron, Y.; Preston, B.D.; Dougherty, J.P. Genetic rearrangements occurring during a single cycle of murine leukemia virus vector replication: characterization and implications. J. Virol. 1995, 69, 7991–8000. [Google Scholar] [PubMed]
  208. Panganiban, A.T.; Fiore, D. Ordered interstrand and intrastrand DNA transfer during reverse transcription. Science 1988, 241, 1064–1069. [Google Scholar] [CrossRef] [PubMed]
  209. Koo, H.M.; Brown, A.M.; Ron, Y.; Dougherty, J.P. Spleen necrosis virus, an avian retrovirus, can infect primate cells. J. Virol. 1991, 65, 4769–4776. [Google Scholar] [PubMed]
  210. Chiu, C.N.; Mitra, R.; Chiu, I.-M. Exchange of genetic sequences of long terminal repeat and the env gene by a promiscuous primate type D retrovirus. Virus Res. 2003, 96, 107–111. [Google Scholar] [CrossRef]
  211. Wayengera, M. On the general theory of the origins of retroviruses. Theor. Biol. Med. Model. 2010, 7, 5. [Google Scholar] [CrossRef] [PubMed]
  212. Nikolaitchik, O.A.; Galli, A.; Moore, M.D.; Pathak, V.K.; Hu, W.-S. Multiple barriers to recombination between divergent HIV-1 variants revealed by a dual-marker recombination assay. J. Mol. Biol. 2011, 407, 521–531. [Google Scholar] [CrossRef] [PubMed]
  213. Galli, A.; Kearney, M.; Nikolaitchik, O.A.; Yu, S.; Chin, M.P.S.; Maldarelli, F.; Coffin, J.M.; Pathak, V.K.; Hu, W.-S. Patterns of Human Immunodeficiency Virus type 1 recombination ex vivo provide evidence for coadaptation of distant sites, resulting in purifying selection for intersubtype recombinants during replication. J. Virol. 2010, 84, 7651–7661. [Google Scholar] [CrossRef] [PubMed]
  214. Rhodes, T.; Wargo, H.; Hu, W.-S. High rates of human immunodeficiency virus type 1 recombination: near-random segregation of markers one kilobase apart in one round of viral replication. J. Virol. 2003, 77, 11193–11200. [Google Scholar] [CrossRef] [PubMed]
  215. Flynn, J.A.; An, W.; King, S.R.; Telesnitsky, A. Nonrandom dimerization of murine leukemia virus genomic RNAs. J. Virol. 2004, 78, 12129–12139. [Google Scholar] [CrossRef] [PubMed]
  216. Roda, R.H.; Balakrishnan, M.; Hanson, M.N.; Wöhrl, B.M.; Le Grice, S.F.J.; Roques, B.P.; Gorelick, R.J.; Bambara, R.A. Role of the Reverse Transcriptase, Nucleocapsid Protein, and Template Structure in the Two-step Transfer Mechanism in Retroviral Recombination. J. Biol. Chem. 2003, 278, 31536–31546. [Google Scholar] [CrossRef] [PubMed]
  217. Moumen, A.; Polomack, L.; Roques, B.; Buc, H.; Negroni, M. The HIV-1 repeated sequence R as a robust hot-spot for copy-choice recombination. Nucleic Acids Res. 2001, 29, 3814–3821. [Google Scholar] [CrossRef] [PubMed]
  218. Dykes, C.; Balakrishnan, M.; Planelles, V.; Zhu, Y.; Bambara, R.A.; Demeter, L.M. Identification of a preferred region for recombination and mutation in HIV-1 gag. Virology 2004, 326, 262–279. [Google Scholar] [CrossRef] [PubMed]
  219. Galli, A.; Lai, A.; Corvasce, S.; Saladini, F.; Riva, C.; Dehò, L.; Caramma, I.; Franzetti, M.; Romano, L.; Galli, M.; et al. Recombination analysis and structure prediction show correlation between breakpoint clusters and RNA hairpins in the pol gene of human immunodeficiency virus type 1 unique recombinant forms. J. Gen. Virol. 2008, 89, 3119–3125. [Google Scholar] [CrossRef] [PubMed]
  220. Galetto, R.; Moumen, A.; Giacomoni, V.; Véron, M.; Charneau, P.; Negroni, M. The structure of HIV-1 genomic RNA in the gp120 gene determines a recombination hot spot in vivo. J. Biol. Chem. 2004, 279, 36625–36632. [Google Scholar] [CrossRef] [PubMed]
  221. Galetto, R.; Giacomoni, V.; Véron, M.; Negroni, M. Dissection of a circumscribed recombination hot spot in HIV-1 after a single infectious cycle. J. Biol. Chem. 2006, 281, 2711–2720. [Google Scholar] [CrossRef] [PubMed]
  222. Certo, J.L.; Kabdulov, T.O.; Paulson, M.L.; Anderson, J.A.; Hu, W.S. The nucleocapsid domain is responsible for the ability of spleen necrosis virus (SNV) Gag polyprotein to package both SNV and murine leukemia virus RNA. J. Virol. 1999, 73, 9170–9177. [Google Scholar] [PubMed]
  223. Méric, C.; Goff, S.P. Characterization of Moloney murine leukemia virus mutants with single-amino-acid substitutions in the Cys-His box of the nucleocapsid protein. J. Virol. 1989, 63, 1558–1568. [Google Scholar] [PubMed]
  224. Rein, A.; Henderson, L.E.; Levin, J.G. Nucleic-acid-chaperone activity of retroviral nucleocapsid proteins: significance for viral replication. Trends Biochem. Sci. 1998, 23, 297–301. [Google Scholar] [CrossRef]
  225. Zhang, Y.; Qian, H.; Love, Z.; Barklis, E. Analysis of the assembly function of the human immunodeficiency virus type 1 gag protein nucleocapsid domain. J. Virol. 1998, 72, 1782–1789. [Google Scholar] [PubMed]
  226. Darlix, J.L.; Lapadat-Tapolsky, M.; de Rocquigny, H.; Roques, B.P. First glimpses at structure-function relationships of the nucleocapsid protein of retroviruses. J. Mol. Biol. 1995, 254, 523–537. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The primary packaging determinants of retroviruses are located at the 5’ end of the viral genome, within the 5’ untranslated region (UTR) and 5’ end of gag. (A) Schematic representation of the location of the packaging determinants within the 5’ UTR and gag for different retroviruses; (B) Generic representation of the higher order structures (stem loops) predicted, biochemically validated, or both at the 5’ end of different retroviruses important for gRNA dimerization and encapsidation. For the sake of simplicity and distinction from long terminal repeats (LTR) sequences, the 5’ UTR in this figure and Table 1 refers to sequences starting from the end of 5’ LTR (and not R) to the beginning of gag gene (excluding the ATG). Figure not drawn to scale. HIV: human immunodeficiency virus; SIV: simian immunodeficiency virus; FIV: feline immunodeficiency virus; MPMV: Mason-Pfizer monkey virus; MMTV: mouse mammary tumor virus; BLV: bovine leukemia virus; HTLV-1: human T-lymphotropic virus type 1; MLV: murine leukemia virus; SNV: avian spleen necrosis virus; RSV: Rous sarcoma virus; HFV: human foamy viruses.
Figure 1. The primary packaging determinants of retroviruses are located at the 5’ end of the viral genome, within the 5’ untranslated region (UTR) and 5’ end of gag. (A) Schematic representation of the location of the packaging determinants within the 5’ UTR and gag for different retroviruses; (B) Generic representation of the higher order structures (stem loops) predicted, biochemically validated, or both at the 5’ end of different retroviruses important for gRNA dimerization and encapsidation. For the sake of simplicity and distinction from long terminal repeats (LTR) sequences, the 5’ UTR in this figure and Table 1 refers to sequences starting from the end of 5’ LTR (and not R) to the beginning of gag gene (excluding the ATG). Figure not drawn to scale. HIV: human immunodeficiency virus; SIV: simian immunodeficiency virus; FIV: feline immunodeficiency virus; MPMV: Mason-Pfizer monkey virus; MMTV: mouse mammary tumor virus; BLV: bovine leukemia virus; HTLV-1: human T-lymphotropic virus type 1; MLV: murine leukemia virus; SNV: avian spleen necrosis virus; RSV: Rous sarcoma virus; HFV: human foamy viruses.
Viruses 08 00276 g001
Figure 2. Multiple sequence alignment of the nucleocapsid region of the various Gag proteins of the listed retroviruses. The zinc finger binding domains present in the protein are boxed. The conserved cysteine (C) and histidine (H) amino acids of the zinc finger domains are shown in red. As can be observed, MLV and SNV contain only one zinc finger binding domain. The spumaviruses (e.g., HFV) are unique among retroviruses in that they contain three glycine-arginine-rich boxes instead of the cysteine-histidine-containing zinc finger domains [32]. Dashes have been added to facilitate the alignment of the differently sized nucleocapsid proteins.
Figure 2. Multiple sequence alignment of the nucleocapsid region of the various Gag proteins of the listed retroviruses. The zinc finger binding domains present in the protein are boxed. The conserved cysteine (C) and histidine (H) amino acids of the zinc finger domains are shown in red. As can be observed, MLV and SNV contain only one zinc finger binding domain. The spumaviruses (e.g., HFV) are unique among retroviruses in that they contain three glycine-arginine-rich boxes instead of the cysteine-histidine-containing zinc finger domains [32]. Dashes have been added to facilitate the alignment of the differently sized nucleocapsid proteins.
Viruses 08 00276 g002
Figure 3. Comparison of the major cis-acting sequences among different retroviruses required for the reverse transcription and integration steps of retroviral life cycle. The canonical TG and CA dinucleotides within the 3’ and 5’ attachment (att) sequences, respectively, are boxed. All sequences are compared to either that of HIV-1 or MPMV, where a “-“ represents homology, while differences are shown by the native nucleotide sequence. PBS: primer binding site; PPT: polypurine tract; U3 att: 3’ attachment site in the U3 region; U5 att: 5’ attachment site in the U5 region of the long terminal repeat.
Figure 3. Comparison of the major cis-acting sequences among different retroviruses required for the reverse transcription and integration steps of retroviral life cycle. The canonical TG and CA dinucleotides within the 3’ and 5’ attachment (att) sequences, respectively, are boxed. All sequences are compared to either that of HIV-1 or MPMV, where a “-“ represents homology, while differences are shown by the native nucleotide sequence. PBS: primer binding site; PPT: polypurine tract; U3 att: 3’ attachment site in the U3 region; U5 att: 5’ attachment site in the U5 region of the long terminal repeat.
Viruses 08 00276 g003
Table 1. Summary of the packaging determinants mapped to the R/U5, 5’ untranslated region (5’ UTR)* and gag gene of different retroviruses.
Table 1. Summary of the packaging determinants mapped to the R/U5, 5’ untranslated region (5’ UTR)* and gag gene of different retroviruses.
VirusLength of 5’ UTR*5’ UTR* and R/U5 Sequences Requiredgag Sequences RequiredReference
HIV-1~152 ntsEntire UTR, SL1 and SL3<455 nts[18,29]
HIV-2~246 nts140 nts366 nts[40,41,42,43]
SIVmac239~231 nts117–175 nts upstream of gag ATG (SL1, SL3 and SL4)?[29,44,45]
FIV~270 ntsFirst 150 nts100 nts[46]
MPMV~128 ntsFirst 50 nts + last 23 nts100 nts[47,48]
MMTV~160 ntsEntire UTR120 nts[49]
BLV~94 ntsEntire UTRFirst 70 nts of gag (SL1 and SL2) + 132 nts (nt 1015–1147) in CA[50,51,52]
HTLV-1~94 nts?~70 nts in gag (SL1 and SL2)[53]
MLV~476 nts350 nts (nt 215–565) of 5’ UTR328 nts [54,55,56]
SNV (REV)~401 nts180 nts (SL1 and SL2)?[57,58]
RSV~278 nts82 nts (O3 stem)?[59,60,61]
HFV~100 ntsEntire UTR?[34,35,36]
5’ UTR: 5’untranslated region here is defined as the region between the start of the primer binding site and before gag ATG. HIV: human immunodeficiency virus; SIV: simian immunodeficiency virus; FIV: feline immunodeficiency virus; MPMV: Mason-Pfizer monkey virus; MMTV: mouse mammary tumor virus; BLV: bovine leukemia virus; HTLV-1: human T-lymphotropic virus type 1; MLV: murine leukemia virus; SNV: avian spleen necrosis virus; REV: Reticuloendothelial virus; RSV: Rous sarcoma virus; HFV: human foamy viruses; SL: stem loop; CA: capsid.
Table 2. Summary of RNA cross-/co-packaging and propagation among retroviruses.
Table 2. Summary of RNA cross-/co-packaging and propagation among retroviruses.
ProteinHIV-1HIV-2SIVFIVMPMVMMTVSNV/REVMLVRSVBLV
RNAPACKPROPPACKPROPPACKPROPPACKPROPPACKPROPPACKPROPPACKPROPPACKPROPPACKPROPPACKPROP
HIV-1 [40,137][40,137]ND[30][161][161][24][24][125]*ND[28][28][110][30]
HIV-2[40,137][40,137] [137][137]
SIV[29,137][29,137][137][137] [161][161][24][24]
FIV[161][161] [161][161] [24][24]
MPMV[24][24] [24][24][161][161] [25][25]
MMTV[125]ND [25][25] [163][163]
SNV[28][28] [138][138]
MLV[110]*ND ND[30] [31,58], [121]*, [138,141][31,58], [121]*, [138,141] [119]*[119]*[52]ND
RSV
BLV
Green: Yes; Red: No; *: RNA packaging tested in a chimeric Gag context; Grey: RNA packaging and propagation using proteins from the same virus; ND: not done.

Share and Cite

MDPI and ACS Style

Ali, L.M.; Rizvi, T.A.; Mustafa, F. Cross- and Co-Packaging of Retroviral RNAs and Their Consequences. Viruses 2016, 8, 276. https://0-doi-org.brum.beds.ac.uk/10.3390/v8100276

AMA Style

Ali LM, Rizvi TA, Mustafa F. Cross- and Co-Packaging of Retroviral RNAs and Their Consequences. Viruses. 2016; 8(10):276. https://0-doi-org.brum.beds.ac.uk/10.3390/v8100276

Chicago/Turabian Style

Ali, Lizna M., Tahir A. Rizvi, and Farah Mustafa. 2016. "Cross- and Co-Packaging of Retroviral RNAs and Their Consequences" Viruses 8, no. 10: 276. https://0-doi-org.brum.beds.ac.uk/10.3390/v8100276

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop