ArticlePDF AvailableLiterature Review

Requirements for efficient minus strand strong-stop DNA transfer in human immunodeficiency virus 1

Taylor & Francis
RNA Biology
Authors:

Abstract and Figures

After HIV-1 enters a human cell, its RNA genome is converted into double stranded DNA during the multistep process of reverse transcription. First (minus) strand DNA synthesis is initiated near the 5' end of the viral RNA, where only a short fragment of the genome is copied. In order to continue DNA synthesis the virus employs a complicated mechanism, which enables transferring of the growing minus strand DNA to a remote position at the genomic 3' end. This is called minus strand DNA transfer. The transfer enables regeneration of long terminal repeat sequences, which are crucial for viral genomic DNA integration into the host chromosome. Numerous factors have been identified that stimulate minus strand DNA transfer. In this review we focus on describing protein-RNA and RNA-RNA interactions, as well as RNA structural features, known to facilitate this step in reverse transcription.
Content may be subject to copyright.
REVIEW
RNA Biology 8:2, 230-236; March/April 2011; © 2011 Landes Bioscience
230 RNA Biology Volume 8 Issue 2
The human immunodeficiency virus type 1 (HIV-1) belongs to
the retrovirus family, members of which have two copies of a
single stranded RNA genome. The genome encodes only a small
set of genes important for viral survival, relying primarily on
host cell factors to complete the viral life cycle. As with all ret-
roviruses, the replication of HIV-1 requires the RNA genome to
be reverse transcribed into DNA soon after the virus enters the
cell. The DNA copy of the virus is subsequently transported into
nucleus for incorporation into human DNA to become perma-
nently linked with its host. The DNA form of HIV-1 or provirus,
may stay latent for a long period of time, but eventually will be
expressed to form new viral particles, which destroy the cell and
go on to infect new cells.
Reverse transcription has been a main target in treatment of
HIV infections, as it is an imperative step in viral replication.
Extensive research is being conducted to understand the mecha-
nisms of this multistep process. Drugs developed to interfere
*Correspondence to: Robert A. Bambara; Email: robert_bambara@urmc.
rochester.edu
Submitted: 10/01/10; Revised: 01/05/11; Accepted: 01/11/11
DOI: 10.4161/rna.8.2.14802
After HIV-1 enters a human cell, its RNA genome is converted
into double stranded DNA during the multistep process of
reverse transcription. First (minus) strand DNA synthesis is
initiated near the 5' end of the viral RNA, where only a short
fragment of the genome is copied. In order to continue DNA
synthesis the virus employs a complicated mechanism, which
enables transferring of the growing minus strand DNA to a
remote position at the genomic 3' end. This is called minus
strand DNA transfer. The transfer enables regeneration of long
terminal repeat sequences, which are crucial for viral genomic
DNA integration into the host chromosome. Numerous factors
have been identied that stimulate minus strand DNA transfer.
In this review we focus on describing protein-RNA and RNA-
RNA interactions, as well as RNA structural features, known to
facilitate this step in reverse transcription.
Requirements for ecient minus strand
strong-stop DNA transfer in human
immunodeciency virus 1
Dorota Piekna-Przybylska and Robert A. Bambara*
Depart ment of Biochemistr y and Biophysics and the Ce nter for RNA Biology; Un iversity of Roches ter School of Medicine a nd Dentistry; Roc hester, NY USA
Key words: minus strand DNA transfer, reverse transcription, RNA genome circularization, HIV-1, retroviruses
Abbreviations: LTR, long terminal repeat; PBS, primer binding site; (-)ssDNA, minus strong stop DNA; RT, reverse transcriptase;
NC, nucleocapsid protein; DIS, dimerization initiation site; TAR, transacting responsive hairpin
with viral reverse transcription have provided effective treatment,
because they prevent the completion of proviral DNA synthesis
and its integration in the host genome. In this review, we focus on
an early step in HIV-1 reverse transcription, called minus strand
DNA transfer. This is a complicated, but obligatory event in con-
version of the RNA genome into DNA and critical in generation of
the DNA provirus in a form for its integration into host cells DNA.
First (minus) Strand DNA Transfer
The genomic RNA of HIV-1 consists of several major genes that
code for essential proteins and enzymes needed for viral replica-
tion and maturation. The protein coding region is flanked by two
unique sequences, U5, at the 5' end and U3 at the 3' end and two
identical sequences of the repeat (R) elements at both ends. These
regions do not encode proteins, but contain regulatory sequences
important for viral replication. During the process of reverse
transcription, U5 and U3 are duplicated to create U3-R-U5 at
both ends of the proviral DNA, segments known as long terminal
repeats (LTRs). These are important regions used by the virus to
mediate its own integration into the host chromosome. In addi-
tion, the upstream LTR acts as a promoter and enhancer and the
downstream LTR acts as transcription termination and polyad-
enylation site.1 The binding of cellular and viral proteins to these
regions regulates HIV gene expression.2 ,3
Duplication of unique sequences U3 and U5 in the LTRs is a
result of the minus strand DNA transfer during reverse transcrip-
tion (Fig. 1A). Conversion of RNA into DNA is performed by
the viral enzyme reverse transcriptase (RT). The process of first
(minus) strand DNA synthesis is initiated from a host cellular
tRNA3Lys that is captured within the viral particles when they
are formed. The 18 nt at the 3' end of the tRNA are bound to
the complementary sequence of the primer binding site (PBS)
localized 182 nt from the 5' end of the HIV-1 RNA genome, just
upstream of the unique sequence U5. The RT extends the primer-
tRNA copying the U5 and R elements and synthesizing 181 nt
long cDNA known as a minus strong stop DNA ((-)ssDNA).
The synthesis of minus strand DNA can be continued only after
the (-)ssDNA is transferred to the 3' end of viral RNA genome
www.landesbioscience.com RNA Biology 231
REVIEW REVIEW
different RNA-RNA and RNA-protein interactions have evolved
to ensure that it is rapid and efficient.5,6
Protein Activities Promoting Minus Strand Transfer
Conversion of genomic RNA into the DNA provirus requires two
proteins encoded by the genome, the RT and nucleocapsid pro-
tein (NC). In HIV-1, the polymerization activity of RT is accom-
panied by an R Nase H activity. RNase H is essential for RNA
degradation within the hybrid duplexes of synthesized cDNA and
the RNA that was used as a template during reverse transcription.
The RT is composed of two subunits, p66 and p51 (66 and 51
kDa). Both activities are located in subunit p66, whereas p51 acts
as a structural polypeptide for proper conformation of p66.7, 8 The
p51 is formed by proteolytic cleavage of the C-terminal domain in
p66 subunit.9 The RNase H active site in RT is located in the C
domain and is separated from the polymerization site by 18 bp in
DNA:RNA heteroduplex substrates.10,11 This configuration allows
RT to make cleavages in RNA during cDNA synthesis, which is
known as polymerization-dependent RNase H activity.12 However,
the polymerization rate of RT is greater than the rate of RNA
hydrolysis, thus polymerization-independent cuts, made during
revisits of RT molecules to remaining RNA oligonucleotides, are
necessary for complete removal of the genomic template.12,13 A sin-
gle virion contains about 50 molecules of RT enzyme.14 The excess
RT molecules, beyond what is apparently required for synthesis,
may be important in carrying out RNase H cleavage in order to
clear the minus strand DNA to allow efficient strand transfer,
accurate priming of plus strand synthesis and unimpeded synthesis
of the plus strand.
Secondary structures in the RNA genome, such as hairpins
and G-quartets can pause the RT during DNA synthesis pro-
moting RNase H cleavages in the RNA template.15,1 6 Studies in
vitro have demonstrated that the TAR hairpin at the 5' end of
the RNA genome causes a major pause in synthesis by RT.17 The
associated RNase H activity of RT cleaves RNA approximately
14–20 nucleotides downstream from the pause site within the
polyA hairpin, that serves as an initiation site for the invasion-
driven mechanism of minus strand transfer (Fig. 1B). The RNase
H cuts create a short gap where the homologous sequence of R
from the 3' end can invade and anneal to the cDNA and initi-
ate displacement of adjacent segments of the 5' end RNA. The
strand exchange continues by a branch migration process until
it reaches the 3' terminus of synthesized (-) ssDNA completing
minus strand DNA transfer.18,19
The polymerization activity of RT enzyme is greatly facili-
tated by NC protein. In fact, many steps of the retroviral life
cycle, including the assembly of virus particles, genomic RNA
dimerization and packaging, reverse transcription and integra-
tion into the host genome require NC involvement.20 -24 NC is a
small (55 amino acid) protein with nucleic acid binding activity,
processed from a precursor polypeptide encoded by the gag gene.
A single virion contains 2,000–3,000 molecules of NC, coating
the RNA genome with binding sites of 7–8 nt.25,26 The key func-
tion of NC is chaperone activity, which is refolding of nucleic
acids into the most thermodynamically stable conformations that
and this process is known as a minus strand DNA transfer. Since
both R elements in the viral genome are identical, the (-)ssDNA
will interact with the 3' end of the RNA genome through comple-
mentarity of their sequences. The consequence of minus strand
transfer is that the U3 region can be copied and the first U3-R-U5
sequence of the 3' LTR is created. Subsequently, this region is used
as a template to synthesize a second (plus) strand DNA, which is
next transferred to the beginning of the viral genome (plus strand
transfer) in order to create the 5' LTR (see Fig. 1 in ref. 4).
Since replication at the ends of the genome is not continuous
and requires interaction between distant related sequences of R at
the 3' end and in growing cDNA at the 5' end, a number of fac-
tors are involved to ensure effective reverse transcription, which is
necessary for viral fitness and long term survival. The first strand
transfer is a rate-limiting step in reverse transcription and many
Figure 1. Minus strand DNA transfer in HIV-1. (A) After entering a cell,
HIV-1 undergoes a process of conversion of its RNA genome into double
stranded DNA. The DNA synthesis is initiated by the extension of the 3'
end of tRNA3Ly s (red) used as a primer to start reverse transcription. The
U5 and R regions are copied into the minus strong stop DNA ((-)ssDNA)
which is transferred to the 3' end of the RNA genome during the pro-
cess of minus strand transfer. The nal product of reverse transcription
is the DNA provirus with two LTRs anking the ends of the HIV genome.
(B) In the invasion-driven mechanism of transfer, the pausing of RT (blue
oval) at the base of the TAR (hairpin with green loop) initiates RNase
H cleavages (blue triangle) within the polyA hairpin (orange loop)
creating a gap for the invasion of the 3' end of the viral genome and
interaction with the nascent cDNA. This initiates the strand exchange
and displacement of adjacent segments of the 5' end RNA until the 3'
terminus of the synthesized (-)ssDNA is fully transferred (not shown).
232 RNA Biology Volume 8 Issue 2
The TAR hairpin has the same structure in LDI and BMH.
Interestingly, structure analyses of both RNA donors, D199 and
D520, revealed that each adopts a different conformation. The
longer RNA donor has a structure similar to LDI, whereas the
shorter RNA donor with lower efficiency of minus strand transfer
resembled the 5'-side of the BMH.40 Possibly the conversion of
one structure to the other occurs during (-)ssDNA synthesis and
transfer, and aids the process. Interestingly, the BMH structure
was proposed to favor RNA genome dimerization and pack-
aging into the virion.42-44 With the DIS formed as a hairpin
in BMH conformation, the kissing loop base pairing between
have the maximal number of base pairs.27 The protein has two
zinc fingers for interaction with ssR NA, ssDNA and dsDNA,
and has the ability to destabilize nucleic acid helices and cause
nucleic acid aggregation.2 8,29
Studies in vitro have shown that the presence of NC during
reverse transcription increases the efficiency of the various steps
and reactions. NC transiently eliminates secondary structures
such as hairpins.30-33 This activity of NC greatly reduces RT paus-
ing and increases the efficiency of DNA synthesis. During syn-
thesis of minus strand DNA the highly structured TAR hairpin
at the 5' end of the RNA template is destabilized with the help
of NC. Although, the pausing of RT is reduced at TAR, the effi-
ciency of minus strand transfer is higher in the presence of NC.
Studies by Purohit and coworkers showed that while the pausing
of RT diminishes in the presence of NC protein, some RNase H
cleavages increase due to enhanced annealing of cDNA to the
RNA template.34 Moreover, NC enhances annealing between
nascent DNA and the invading 3' R sequence and subsequently
promotes strand exchange, which progresses continuously until
the minus strand transfer process is completed.18,19,29,35 ,36
Analyses of minus strand transfer in vitro have demon-
strated that NC protein also significantly inhibits a self-priming
effect.37, 3 8 The 3' end of (-)ssDNA corresponds to the sequence
of TAR, and so this region has the potential to fold back and
form a similar hairpin, which can self-prime DNA synthesis and
inhibit minus strand transfer. The primary basis of self-priming
suppression in the presence of NC is promotion of an exchange
of the very 5'-most RNA oligomer left from polymerization-
dependent RNase H with the homologous RNA sequence of
the genomic 3' end, leading to minus strand DNA transfer.39
Local RNA Structure as an Important Factor
in Minus Strand Transfer
The reconstituted systems used to analyze minus strand DNA
transfer in vitro have demonstrated that using different lengths of
the RNA representing the 5' end of HIV-1 (donor RNA) results in
different transfer efficiencies. The cDNA synthesized in vitro on
the RNA template spanning the region from the 5' end up to PBS
(D199) transfers with low efficiency to the second RNA (acceptor
RNA) representing the 3' end of the virus. However, the efficiency
of transfer increases dramatically when donor RNA is extended at
its 3' end (D520) to include naturally occurring sequences present
beyond the PBS.40 The transfer of cDNA synthesized from both
donor RNAs uses the same acceptor invasion-driven mechanism,
but that mechanism is more effective when a longer donor RNA is
used. This implies that folding properties of donor RNAs having
different lengths influences the transfer reaction, indicating that
local structure is an important influence on minus strand transfer.40
Analyses in vitro demonstrated that the 5'-untranslated region
in HIV-1 can adopt two distinct structures (Fig. 2A).41 A long-
distance base-pairing interaction (LDI) between the polyA and
dimerization initiation site (DIS) can be formed into a ther-
modynamically stable structure. However, the sequence can
refold into the branched multiple hairpin (BMH) conformation,
which allows the polyA and DIS motifs to fold into hairpins.
Figure 2. The local structure of the 5' end of the HIV-1 RNA genome
and its proposed interactions with the 3' end. (A) The untranslated
RNA at the 5' end can adopt two alternative structures: a long distance
interaction (LDI) and a branched multiple-hairpin (BMH) structure.
Sequences of loops of the hairpins TAR (green, in HIV-1 positions
1–57), polyA (orange, 66–95), PBS (black, 182–199), DIS (red, 243–277),
SD (blue, 282–300), Ψ (yellow, 306–331) are indicated. (B) Proposed
RNA-RNA and cDNA-RNA interactions during circularization of the RNA
genome in HIV-1. Sequences of loops of the hairpins TAR (green, 1–57
and 9,075–9,135), polyA (orange, 66–95 and 9,141–9,170) and PBS (black,
182–199) are indicated. Sequences of motif 9 nt (light blue, 9,033–9,041),
segment 1 (dark blue, 9,008–9,019), gag (619–696), tRNA (red) and the
cDNA (grey) are also shown. The bulk of the genome is designated by a
dotted line. Potential regions of interaction are shown in proximity.
www.landesbioscience.com RNA Biology 233
indicated in retroviruses 35 years ago.56,57 In subsequent years it was
discovered that many single-stranded RNA viruses adopt a circular
conformation, which is essential for different stages of their viral
life cycles.58 Viral genome circularization is known to stimulate
initiation of translation, as is the case for many cellular mRNAs.58
Juxtaposing of viral ends also facilitates transcription, genome rep-
lication and viral packaging.59-62 In many positive strand viruses,
such as ratoviruses, picornaviruses and pestiviruses, a 5'-3' end con-
tact is mediated through RNA binding protein interactions, which
can involve both viral and cellular factors.60,63-66
Alternatively, viruses can circularize their genomes via long-
distance RNA-RNA interactions. With application of atomic
force microscopy it was shown that the dengue virus genome can
form a circle in the absence of proteins.67 RNA-RNA genomic
cyclizations have been documented in many plant viruses,59 fla-
viviruses,60,68-70 hepatitis C virus,71 FMDV (Foot-and-mouth dis-
ease virus),72 and also in HIV-1.73
Several different types of RNA-RNA interactions were
described that could mediate HIV-1 genome circularization. A
palindromic sequence of 10 nucleotides in stem-loop of the TAR
hairpin in the R element was indicated as a possible place of
interaction involved in genome dimerization, in addition to the
primary site of dimerization at DIS.74 Since the TAR is present
at both ends of the viral RNA, the palindromic sequences could
interact and circularize the HIV-1 genome (Fig. 2B).75 This is a
way that the sequences of R elements involved in minus strand
transfer would self-juxtapose by direct interaction. Stem-loops
and loop-loop kissing structures formed by sequences at both
ends of the RNA genome were demonstrated to participate in
genome circularization in many plant viruses.59 Howe ver, ana l-
yses of mutations in TAR sequences that should have disrupted
the putative binding showed that 3'-5' end interactions in vitro
were not significantly affected.75 Apparently, other factors in
addition to TAR hairpins maintain genome circularization,
whereas putative base pairing between TAR sequences at both
ends of the genome might still help in juxtaposing R elements
for minus strand transfer.
Another type of RNA-RNA interaction involves a region in
gag and the 3' U3/R (Fig. 2B). Here, a structure is formed with
extensive base pairing among nucleotides within the gag ORF
(positions 619–696, clone pNL43) and sequences flanking the
TAR hairpin, 18 nt of U3 (positions 9,054–9,071) and the polyA
hairpin of the 3' R (positions 9,136–9,170).73
Interactions between cis-acting sequences of viral RNA
genomes were initially documented for many flaviviruses.61 ,67-7 0
Here, genome circularization is maintained by short comple-
mentary cyclization sequences (CS) present in the capsid coding
region and at the 3' end of the RNA genome. For dengue virus,
additional sequences described as UARs (upstream AUG regions)
were also found in proximity to the CS, and UARs were shown
to be significant for viral viability.67 A similar 3'-5' RNA-RNA
interaction was reported for yeast LTR Ty1 retrotransposon.76
LTR retrotransposons are thought to be ancestors of retrovi-
ruses.7 7,7 8 The interactions involve base pairing of complementary
14 nucleotide long sequences, called CYC5 in the gag ORF and
CYC3 in U3.76 The reconstituted system in vitro demonstrated
complementary sequences in the loop will lead to dimerization
of viral genomes.45,46 These structural inter-conversions may
coordinate sequential events in virus assembly, reverse tran-
scription and protein translation.
Extensive studies in vitro revealed that the structure of the
sequences involved directly in the minus strand transfer is also
important. The sizes of the R elements in different retroviruses
vary significantly from 16 bases in mouse mammary tumor virus
(MMTV) to 247 bases in human T-cell leukemia virus type 2.4 7,4 8
However, a significant shortening of the R sequences affects the
efficiency of minus strand transfer.49-51 Moreover, early strand
transfers of partially synthesized cDNA are not frequently seen
wit h HI V-1.52 Although the (-)ssDNA synthesized at the 5' end
of the viral genome has to hybridize with the complementary R
region at the 3' end, the length of complementarity is apparently
not the only factor important for efficient minus strand transfer.
The structure of the R elements is also crucial. A stable TAR hair-
pin at the 5' end is important for efficient minus strand transfer,
but disruption of this hairpin at the 3' end of viral genome is ben-
eficial for the transfer in vitro. However, stabilizing the polyA hair-
pin in the 5' and 3' R will inhibit the transfer reaction.53 Studies in
vitro have demonstrated that pre-incubation of NC with the 3' end
RNA template rather than with just the 5' end RNA, significantly
enhances the efficiency of minus strand transfer. This implies that
destabilization of hairpins at the 3' end by NC is beneficial for the
reaction.54 Overall, strong structure in the genomic 3' end appears
to decrease the efficiency of the (-)ssDNA transfer, but a highly
structured 5' end of the genome promotes the transfer.
The stimulatory effect of the folded genomic 5' end on minus
strand transfer is presumably important for the invasion-driven
mechanism of transfer, as described above (Fig. 1B). TAR struc-
ture pauses RT during cDNA synthesis, promoting R Nase H
cleavages in the copied template creating an invasion site for the
3' R RNA.17 Lack of structure in the 3' RNA likely promotes its
ability to invade and base pair with the cDNA.
Proximity of R Elements
in Minus Strand DNA Transfer
Analyses in vitro and in vivo have demonstrated that the length
of homology between the 97 nucleotide R elements in HIV-1
influences the efficiency of minus strand transfer.50,55 However, it
was shown that transfer within a homology of just 20 nucleotides
will not be affected in vitro, as long as the mutated 3' R sequence
still contains an unchanged polyA hairpin, needed for invasion
of the (-)ssDNA.55 The interaction between the invasion site in
the (-)ssDNA and the invading polyA hairpin of the 3' R brings
sequences involved in the completion of transfer at the (-)ssDNA
terminus into proximity.
The 5' and 3' R elements are over 9,000 nucleotides apart
in the linear form of the HIV-1 RNA genome. However, RNA
molecules are capable of forming complex tertiary interactions
between related sequences over very long distances. Circularization
of the viral genome in a way that juxtaposes the R elements should
facilitate the process of minus strand DNA transfer. The possibility
of genome circularization during reverse transcription was already
234 RNA Biology Volume 8 Issue 2
in Drosophila melanogaster.85 Here, the primer tRNA2Lys inter-
acts at the PBS and within the region upstream of the PPT and
U3 located about 450 nucleotides from the 3' end of the Gypsy
genomic RNA. The second interaction involves the nucleotides
of the TψC arm and Variable loop in tRNA2Lys.
In order to maintain the proximity of R elements for effi-
cient minus strand transfer, the scenario of the 5'-3' end interac-
tions in HIV-1 genomic RNA could be very complex. Consider
that RNA genome circularization might be already established
by the time of packaging of the viral genome. The circular
form of HIV-1 RNA could be maintained via interactions
between gag and U3/R sequences at the genomic 3' end. The
5' end would form a BMH conformation with DIS and TAR
hairpins exposed. Within the virion, the dimerization of two
RNA genomic molecules via TAR hairpins might no longer be
essential and could be replaced to allow base pairing between
TAR hairpins at the 5' and 3' ends. Proximity of the genomic
ends due to gagU3/R contact will facilitate self-juxtaposing
of R elements via stem-loop kissing base pairing of TAR hair-
pins. Soon after reverse transcription begins, the interactions
between U5 and the anticodon stem in tRNA3Lys are unwound
enabling tR NA to bind with motif 9 nt in U3. This interac-
tion could help to maintain the proximity of R elements dur-
ing minus strand transfer, when contacts of the 3' R with gag
sequence need to be removed. As the synthesis of minus strand
DNA proceeds towards the 5' end of the genome, TAR inter-
actions will be disrupted. Furthermore, minus strand transfer
occurs within the 3' R element, and so the interactions with gag
have to be disrupted too. The binding between motif 9 nt and
tRNA3Lys would maintain circularization and bring segment 1
into proximity with the 3' end of the tRNA primer, facilitat-
ing tRNA displacement from the PBS. After the transfer of (-)
ssDNA the growing cDNA will disrupt all binding between U3
sequences and gag and the tRNA. If accurate, this interpreta-
tion explains why there must be a stepwise series of interactions.
Conclusion
The ability of HIV-1 to recombine frequently and error prone
reverse transcription are survival features of the virus that allow
it to evolve rapidly and avoid inactivation by the human immune
system. Drugs currently used to treat HIV infections target the
virus replication process. Although they can effectively inhibit the
viral replication machinery, HIV-1 is so effective at evolving drug
resistance that new treatments need to be developed. Targeting
the steps in viral replication, which are unique to the virus, offers
opportunities for treating viral infections without disruption of
human cellular function. Minus strand DNA transfer might be
a good target for interference of viral replication, as it is an early
and mandatory process preparing viral genetic material for inte-
gration into the host. Moreover, efficient minus strand transfer is
critical for maximum viral replication fitness.
Acknowledgements
Work from our laboratory described herein was supported by the
US National Institutes of Health research grant GM049573.
that contact between CYC5 and CYC3 is important for effi-
cient minus strand transfer and also to enhance initiation of
reverse transcription. Additionally, CYC5-CYC3 interactions
are required for Ty1 transposition.76 In the case of HIV-1, with
application of a reconstituted system, mutation analysis of a 3'
U3/R-gag interaction showed that this structure also enhances
efficiency of minus strand DNA transfer, in experiments using a
DNA primer to start reverse transcription.75
Genome circularization in retroviruses via tertiary interac-
tions between 5' and 3' genome ends may also involve other
nucleic acid molecules, such as the cDNA made de novo and
tRNA3Lys used by HIV-1 as a primer to start reverse transcription.
Specifically, it was proposed that growing cDNA extended from
the tR NA3Lys primer could interact via kissing loop contact of the
TAR hairpin segment in (-)ssDNA and TAR at 3' RNA genome
(Fig. 2B).53 Also, the invasion mechanism of minus strand trans-
fer itself is a mechanism for genome circularization (Fig. 1B).
Here, the interaction between the polyA hairpin at the 3' end
of the RNA and the invasion site in (-)ssDNA could hold both
of the viral genome ends together. In both cases, the interaction
only becomes possible after some (-)ssDNA synthesis.
The primer tRNA3Lys might also serve as a bridging factor inter-
acting with the 3' end of the viral genomic RNA (Fig. 2B). The
interaction would occur between nine nucleotides in U3 (motif
9 nt, positions 9,033–9,041) adjacent to the 3' R element and
nucleotides of the anticodon stem in tRNA3Ly s (positions 38–46).79
Computational analysis revealed that sequences flanking motif 9
nt also have complementarity to tRNA3Ly s, and the whole region
spanning extensive sequences in U3/R might represent an entire
intron-containing tRNA gene incorporated during evolution
of the HIV-1 genome.80 Analyses in vitro and in vivo have dem-
onstrated that motif 9 nt and complementary sequences in U3
(segment 1, positions 9,008–9,019) stimulate minus strand trans-
fer.7 9-81 Interactions between U3 sequences and tRNA3Lys appear
most likely to be established after initiation of reverse transcrip-
tion. The nucleotides of the anticodon stem in tRNA3Lys are first
involved in an interaction with sequences upstream of the PBS in
U5 in order to form the initiation complex for DNA synthesis.82, 83
In the case of segment 1, nucleotides are complementary to the
3' end of tRNA3Lys, where 18 nucleotides of the tRNA serve as a
primer and base pair directly with the PBS. Here too, the interac-
tions might take place only after the start of reverse transcription.
Whereas the binding between tRNA3Lys and motif 9 nt in
U3 could help in maintaining the R elements in proximity, the
significance of interactions with segment 1 might be different.
For example, segment 1 could help in displacement of the tRNA
from the PBS to allow copying of the 3' end of the tRNA into
DNA in preparation for second strand transfer.
Primer tRNA serving as a bridging factor in genome circular-
ization was also indicated for LTR retrotransposons and endog-
enous LTR retroviruses. The tRNAiMet is used as a primer to
start reverse transcription in Ty3 yeast retrotransposons. Whereas
the 3' end of this tRNA is bound at the PBS, the TψC arm and
D arm are involved in interactions close to the 3' end of the Ty3
RNA causing retro-element genome circularization.84 A similar
situation was described for the endogenous retrovirus Gypsy
www.landesbioscience.com RNA Biology 235
References
1. Guntaka RV. Transcription termination and polyadenyl-
ation in retroviruses. Microbiol Rev 1993; 57:511-21.
2. Pereira LA, Bentley K, Peeters A, Churchill MJ,
Deacon NJ. A compilation of cellular transcription
factor interactions with the HIV-1 LTR promoter.
Nucleic Acids Res 2000; 28:663-8.
3. Ramirez de Arellano E, Soriano V, Alcamil J, Holguin
A. New findings on transcription regulation across dif-
ferent HIV-1 subtypes. AIDS Rev 2006; 8:9-16.
4. Basu VP, Song M, Gao L, Rigby ST, Hanson MN,
Bambara RA. Strand transfer events during HIV-1
reverse transcription. Virus Res 2008; 134:19-38.
5. Telesnitsky A, Goff SP. Strong-stop strand transfer
during reverse transcription. In: Skalka AM, Goff SP,
Eds. Reverse Transcriptase. Cold Spring Harbor, NY:
Cold Spring Harbor Laboratory Press 1993; 49-83.
6. Canard B, Sarfati R, Richardson CC. Binding of RNA
template to a complex of HIV-1 reverse transcriptase/
primer/template. Proc Natl Acad Sci USA 1997;
94:11279-84.
7. Le Grice SF, Naas T, Wohlgensinger B, Schatz O.
Subunit-selective mutagenesis indicates minimal poly-
merase activity in heterodimer-associated p51 HIV-1
reverse transcriptase. EMBO J 1991; 10:3905-11.
8. Hostomsky Z, Hostomska Z, Fu TB, Taylor J. Reverse
transcriptase of human immunodeficiency virus type 1:
functionality of subunits of the heterodimer in DNA
synthesis. J Virol 1992; 66:3179-82.
9. Chandra A, Gerber T, Kaul S, Wolf C, Demirhan I,
Chandra P. Serological relationship between reverse
transcriptases from human T-cell lymphotropic viruses
defined by monoclonal antibodies. Evidence for two
forms of reverse transcriptases in the AIDS-associated
virus, HTLV-III/LAV. FEBS Lett 1986; 200:327-32.
10. Furfine ES, Reardon JE. Reverse transcriptase.
RNase H from the human immunodeficiency virus.
Relationship of the DNA polymerase and RNA hydro-
lysis activities. J Biol Chem 1991; 266:406-12.
11. Gopalakrishnan V, Peliska JA, Benkovic SJ. Human
immunodeficiency virus type 1 reverse transcriptase:
spatial and temporal relationship between the poly-
merase and RNase H activities. Proc Natl Acad Sci
USA 1992; 89:10763-7.
12. Schultz SJ, Champoux JJ. RNase H activity: structure,
specificity and function in reverse transcription. Virus
Res 2008; 134:86-103.
13. DeStefano JJ, Buiser RG, Mallaber LM, Myers TW,
Bambara RA, Fay PJ. Polymerization and RNase H
activities of the reverse transcriptases from avian myelo-
blastosis, human immunodeficiency and Moloney
murine leukemia viruses are functionally uncoupled. J
Biol Chem 1991; 266:7423-31.
14. Julias JG, Ferris AL, Boyer PL, Hughes SH. Replication
of phenotypically mixed human immunodeficiency
virus type 1 virions containing catalytically active and
catalytically inactive reverse transcriptase. J Virol 2001;
75:6537-46.
15. Roda RH, Balakrishnan M, Kim JK, Roques BP, Fay
PJ, Bambara RA. Strand transfer occurs in retroviruses
by a pause-initiated two-step mechanism. J Biol Chem
2002; 277:46900-11.
16. Shen W, Gao L, Balakrishnan M, Bambara RA. A
recombination hot spot in HIV-1 contains guano-
sine runs that can form a G-quartet structure and
promote strand transfer in vitro. J Biol Chem 2009;
284:33883-93.
17. Kim JK, Palaniappan C, Wu W, Fay PJ, Bambara RA.
Evidence for a unique mechanism of strand transfer
from the transactivation response region of HIV-1. J
Biol Chem 1997; 272:16769-77.
18. Chen Y, Balakrishnan M, Roques BP, Bambara RA.
Steps of the acceptor invasion mechanism for HIV-1
minus strand strong stop transfer. J Biol Chem 2003;
278:38368-75.
19. Chen Y, Balakrishnan M, Roques BP, Fay PJ, Bambara
RA. Mechanism of minus strand strong stop transfer
in HIV-1 reverse transcription. J Biol Chem 2003;
278:8006-17.
20. Muriaux D, De Rocquigny H, Roques BP, Paoletti
J. NCp7 activates HIV-1Lai RNA dimerization by
converting a transient loop-loop complex into a stable
dimer. J Biol Chem 1996; 271:33686-92.
21. Guo J, Wu T, Anderson J, Kane BF, Johnson DG,
Gorelick RJ, Henderson LE, Levin JG. Zinc finger
structures in the human immunodeficiency virus type
1 nucleocapsid protein facilitate efficient minus- and
plus-strand transfer. J Virol 2000; 74:8980-8.
22. Gorelick RJ, Gagliardi TD, Bosche WJ, Wiltrout TA,
Coren LV, Chabot DJ, et al. 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.
23. Carteau S, Gorelick RJ, Bushman FD. Coupled inte-
gration of human immunodeficiency virus type 1
cDNA ends by purified integrase in vitro: stimula-
tion by the viral nucleocapsid protein. J Virol 1999;
73:6670-9.
24. Carteau S, Batson SC, Poljak L, Mouscadet JF, de
Rocquigny H, Darlix JL, et al. Human immunodefi-
ciency virus type 1 nucleocapsid protein specifically
stimulates Mg2+-dependent DNA integration in vitro.
J Virol 1997; 71:6225-9.
25. Henderson LE, Bowers MA, Sowder RC, 2nd, Serabyn
SA, Johnson DG, Bess JW Jr, et al. Gag proteins of
the highly replicative MN strain of human immuno-
deficiency virus type 1: posttranslational modifica-
tions, proteolytic processings and complete amino acid
sequences. J Virol 1992; 66:1856-65.
26. Dib-Hajj F, Khan R, Giedroc DP. Retroviral nucleo-
capsid proteins possess potent nucleic acid strand
renaturation activity. Protein Sci 1993; 2:231-43.
27. Levin JG, Guo J, Rouzina I, Musier-Forsyth K. Nucleic
acid chaperone activity of HIV-1 nucleocapsid protein:
critical role in reverse transcription and molecular
mechanism. Prog Nucleic Acid Res Mol Biol 2005;
80:217-86.
28. You JC, McHenry CS. HIV nucleocapsid protein.
Expression in Escherichia coli, purification and charac-
terization. J Biol Chem 1993; 268:16519-27.
29. Lapadat-Tapolsky M, De Rocquigny H, Van Gent D,
Roques B, Plasterk R, Darlix JL. Interactions between
HIV-1 nucleocapsid protein and viral DNA may have
important functions in the viral life cycle. Nucleic
Acids Res 1993; 21:831-9.
30. Rodriguez-Rodriguez L, Tsuchihashi Z, Fuentes GM,
Bambara RA, Fay PJ. Influence of human immunode-
ficiency virus nucleocapsid protein on synthesis and
strand transfer by the reverse transcriptase in vitro. J
Biol Chem 1995; 270:15005-11.
31. Tanchou V, Gabus C, Rogemond V, Darlix JL.
Formation of stable and functional HIV-1 nucleopro-
tein complexes in vitro. J Mol Biol 1995; 252:563-71.
32. Wu W, Henderson LE, Copeland TD, Gorelick RJ,
Bosche WJ, Rein A, Levin JG. Human immunodefi-
ciency virus type 1 nucleocapsid protein reduces reverse
transcriptase pausing at a secondary structure near the
murine leukemia virus polypurine tract. J Virol 1996;
70:7132-42.
33. Ji X, Klarmann GJ, Preston BD. Effect of human
immunodeficiency virus type 1 (HIV-1) nucleocapsid
protein on HIV-1 reverse transcriptase activity in vitro.
Biochemistry 1996; 35:132-43.
34. Purohit V, Roques BP, Kim B, Bambara RA.
Mechanisms that prevent template inactivation by
HIV-1 reverse transcriptase RNase H cleavages. J Biol
Chem 2007; 282:12598-609.
35. Tsuchihashi Z, Brown PO. DNA strand exchange
and selective DNA annealing promoted by the human
immunodeficiency virus type 1 nucleocapsid protein. J
Virol 1994; 68:5863-70.
36. Lapadat-Tapolsky M, Pernelle C, Borie C, Darlix JL.
Analysis of the nucleic acid annealing activities of
nucleocapsid protein from HIV-1. Nucleic Acids Res
1995; 23:2434-41.
37. Guo J, Henderson LE, Bess J, Kane B, Levin JG.
Human immunodeficiency virus type 1 nucleocapsid
protein promotes efficient strand transfer and specific
viral DNA synthesis by inhibiting TAR-dependent self-
priming from minus-strand strong-stop DNA. J Virol
1997; 71:5178-88.
38. Lapadat-Tapolsky M, Gabus C, Rau M, Darlix JL.
Possible roles of HIV-1 nucleocapsid protein in the
specificity of proviral DNA synthesis and in its vari-
ability. J Mol Biol 1997; 268:250-60.
39. Hong MK, Harbron EJ, O’Connor DB, Guo J,
Barbara PF, Levin JG, Musier-Forsyth K. Nucleic acid
conformational changes essential for HIV-1 nucleo-
capsid protein-mediated inhibition of self-priming in
minus-strand transfer. J Mol Biol 2003; 325:1-10.
40. Song M, Balakrishnan M, Chen Y, Roques BP, Bambara
RA. Stimulation of HIV-1 minus strand strong stop
DNA transfer by genomic sequences 3' of the primer
binding site. J Biol Chem 2006; 281:24227-35.
41. Huthoff H, Berkhout B. Two alternating structures of
the HIV-1 leader RNA. RNA 2001; 7:143-57.
42. Paillart JC, Dettenhofer M, Yu XF, Ehresmann C,
Ehresmann B, Marquet R. First snapshots of the HIV-1
RNA structure in infected cells and in virions. J Biol
Chem 2004; 279:48397-403.
43. Berkhout B, Ooms M, Beerens N, Huthoff H, Southern
E, Verhoef K. In vitro evidence that the untranslated
leader of the HIV-1 genome is an RNA checkpoint that
regulates multiple functions through conformational
changes. J Biol Chem 2002; 277:19967-75.
44. Abbink TE, Ooms M, Haasnoot PC, Berkhout B. The
HIV-1 leader RNA conformational switch regulates
RNA dimerization but does not regulate mRNA trans-
lation. Biochemistry 2005; 44:9058-66.
45. Skripkin E, Paillart JC, Marquet R, Ehresmann B,
Ehresmann C. Identification of the primary site of the
human immunodeficiency virus type 1 RNA dimeriza-
tion in vitro. Proc Natl Acad Sci USA 1994; 91:4945-
9.
46. Laughrea M, Jette L. A 19-nucleotide sequence
upstream of the 5' major splice donor is part of the
dimerization domain of human immunodeficiency virus
1 genomic RNA. Biochemistry 1994; 33:13464-74.
47. Moore R, Dixon M, Smith R, Peters G, Dickson C.
Complete nucleotide sequence of a milk-transmitted
mouse mammary tumor virus: two frameshift suppres-
sion events are required for translation of gag and pol.
J Virol 1987; 61:480-90.
48. Sodroski J, Trus M, Perkins D, Patarca R, Wong-
Staal F, Gelmann E, et al. Repetitive structure in
the long-terminal-repeat element of a type II human
T-cell leukemia virus. Proc Natl Acad Sci USA 1984;
81:4617-21.
49. Lobel LI, Goff SP. Reverse transcription of retroviral
genomes: mutations in the terminal repeat sequences. J
Virol 1985; 53:447-55.
50. Berkhout B, van Wamel J, Klaver B. Requirements for
DNA strand transfer during reverse transcription in
mutant HIV-1 virions. J Mol Biol 1995; 252:59-69.
51. Pfeiffer JK, Telesnitsky A. Effects of limiting homology
at the site of intermolecular recombinogenic template
switching during Moloney murine leukemia virus rep-
lication. J Virol 2001; 75:11263-74.
52. Klaver B, Berkhout B. Premature strand transfer by the
HIV-1 reverse transcriptase during strong-stop DNA
synthesis. Nucleic Acids Res 1994; 22:137-44.
53. Berkhout B, Vastenhouw NL, Klasens BI, Huthoff H.
Structural features in the HIV-1 repeat region facilitate
strand transfer during reverse transcription. RNA
2001; 7:1097-114.
54. Negroni M, Buc H. Copy-choice recombination by
reverse transcriptases: reshuffling of genetic markers
mediated by RNA chaperones. Proc Natl Acad Sci USA
2000; 97:6385-90.
236 RNA Biology Volume 8 Issue 2
78. Eickbush TH. Origin and evolutionary relationships
of retroelements. In: Morse SS, Ed. The Evolutionary
Biology of Viruses. New York, NY: Raven Press 1994;
121-57.
79. Brule F, Bec G, Keith G, Le Grice SF, Roques BP,
Ehresmann B, et al. In vitro evidence for the interac-
tion of tRNA(3)(Lys) with U3 during the first strand
transfer of HIV-1 reverse transcription. Nucleic Acids
Res 2000; 28:634-40.
80. Piekna-Przybylska D, DiChiacchio L, Mathews DH,
Bambara RA. A sequence similar to tRNA 3 Lys gene is
embedded in HIV-1 U3-R and promotes minus-strand
transfer. Nat Struct Mol Biol 2010; 17:83-9.
81. Piekna-Przybylska D, Dykes C, Demeter LM, Bambara
RA. Sequences in the U3 region of human immuno-
deficiency virus 1 improve efficiency of minus strand
transfer in infected cells. Virology 2010; 410:368-74.
82. Isel C, Ehresmann C, Keith G, Ehresmann B, Marquet
R. Initiation of reverse transcription of HIV-1: second-
ary structure of the HIV-1 RNA/tRNA(3Lys) (tem-
plate/primer). J Mol Biol 1995; 247:236-50.
83. Isel C, Westhof E, Massire C, Le Grice SF, Ehresmann
B, Ehresmann C, et al. Structural basis for the specific-
ity of the initiation of HIV-1 reverse transcription.
EMBO J 1999; 18:1038-48.
84. Gabus C, Ficheux D, Rau M, Keith G, Sandmeyer
S, Darlix JL. The yeast Ty3 retrotransposon contains
a 5'-3' bipartite primer-binding site and encodes
nucleocapsid protein NCp9 functionally homologous
to HIV-1 NCp7. EMBO J 1998; 17:4873-80.
84. Gabus C, Ivanyi-Nagy R, Depollier J, Bucheton A,
Pelisson A, Darlix JL. Characterization of a nucleo-
capsid-like region and of two distinct primer tRNA2Lys
binding sites in the endogenous retrovirus Gypsy.
Nucleic Acids Res 2006; 34:5764-77.
68. Hahn CS, Hahn YS, Rice CM, Lee E, Dalgarno L,
Strauss EG, Strauss JH. Conserved elements in the 3'
untranslated region of flavivirus RNAs and potential
cyclization sequences. J Mol Biol 1987; 198:33-41.
69. Corver J, Lenches E, Smith K, Robison RA, Sando T,
Strauss EG, Strauss JH. Fine mapping of a cis-acting
sequence element in yellow fever virus RNA that is
required for RNA replication and cyclization. J Virol
2003; 77:2265-70.
70. Shurtleff AC, Beasley DW, Chen JJ, Ni H, Suderman
MT, Wang H, et al. Genetic variation in the 3'
non-coding region of dengue viruses. Virology 2001;
281:75-87.
71. Romero-Lopez C, Berzal-Herranz A. A long-range
RNA-RNA interaction between the 5' and 3' ends of
the HCV genome. RNA 2009; 15:1740-52.
72. Serrano P, Pulido MR, Saiz M, Martinez-Salas E. The
3' end of the foot-and-mouth disease virus genome
establishes two distinct long-range RNA-RNA inter-
actions with the 5' end region. J Gen Virol 2006;
87:3013-22.
73. Ooms M, Abbink TE, Pham C, Berkhout B.
Circularization of the HIV-1 RNA genome. Nucleic
Acids Res 2007; 35:5253-61.
74. Andersen ES, Contera SA, Knudsen B, Damgaard CK,
Besenbacher F, Kjems J. Role of the trans-activation
response element in dimerization of HIV-1 RNA. J
Biol Chem 2004; 279:22243-9.
75. Beerens N, Kjems J. Circularization of the HIV-1
genome facilitates strand transfer during reverse tran-
scription. RNA 2010; 16:1226-35.
76. Cristofari G, Bampi C, Wilhelm M, Wilhelm FX,
Darlix JL. A 5'-3' long-range interaction in Ty1 RNA
controls its reverse transcription and retrotransposition.
EMBO J 2002; 21:4368-79.
77. Eickbush TH, Malik HS. Origins and evolution of
retrotransposons. In: Craig NL, Craigie R, Gellert M,
Lambowitz AM, Eds. Mobile DNA II. Washington,
DC: American Society of Microbiology Press 2002;
1111-4.
55. Song M, Basu VP, Hanson MN, Roques BP, Bambara
RA. Proximity and branch migration mechanisms in
HIV-1 minus strand strong stop DNA transfer. J Biol
Chem 2008; 283:3141-50.
56. Taylor JM, Illmensee R. Site on the RNA of an avian
sarcoma virus at which primer is bound. J Virol 1975;
16:553-8.
57. Collett MS, Faras AJ. Evidence for circularization of
the avian oncornavirus RNA genome during proviral
DNA synthesis from studies of reverse transcription in
vitro. Proc Natl Acad Sci USA 1976; 73:1329-32.
58. Edgil D, Harris E. End-to-end communication in the
modulation of translation by mammalian RNA viruses.
Virus Res 2006; 119:43-51.
59. Miller WA, White KA. Long-distance RNA-RNA
interactions in plant virus gene expression and replica-
tion. Annu Rev Phytopathol 2006; 44:447-67.
60. Herold J, Andino R. Poliovirus RNA replication
requires genome circularization through a protein-
protein bridge. Mol Cell 2001; 7:581-91.
61. Khromykh AA, Meka H, Guyatt KJ, Westaway EG.
Essential role of cyclization sequences in flavivirus RNA
replication. J Virol 2001; 75:6719-28.
62. Darlix JL. Control of Rous sarcoma virus RNA trans-
lation and packaging by the 5' and 3' untranslated
sequences. J Mol Biol 1986; 189:421-34.
63. Piron M, Vende P, Cohen J, Poncet D. Rotavirus
RNA-binding protein NSP3 interacts with eIF4GI and
evicts the poly(A) binding protein from eIF4F. EMBO
J 1998; 17:5811-21.
64. Vende P, Piron M, Castagne N, Poncet D. Efficient
translation of rotavirus mRNA requires simultaneous
interaction of NSP3 with the eukaryotic translation
initiation factor eIF4G and the mRNA 3' end. J Virol
2000; 74:7064-71.
65. Isken O, Grassmann CW, Yu H, Behrens SE. Complex
signals in the genomic 3' nontranslated region of
bovine viral diarrhea virus coordinate translation and
replication of the viral RNA. RNA 2004; 10:1637-52.
66. Isken O, Grassmann CW, Sarisky RT, Kann M,
Zhang S, Grosse F, et al. Members of the NF90/NFAR
protein group are involved in the life cycle of a positive-
strand RNA virus. EMBO J 2003; 22:5655-65.
67. Alvarez DE, Lodeiro MF, Luduena SJ, Pietrasanta LI,
Gamarnik AV. Long-range RNA-RNA interactions
circularize the dengue virus genome. J Virol 2005;
79:6631-43.
... This activity is critical for ensuring specific and efficient reverse transcription, including initiation [38,39], as well as the minus-and plus-strand transfer reactions [18,19,37]. For example, in the minus-strand transfer step, NC facilitates annealing of the complementary repeat regions (R, r), which contain the highly structured transactivation response element (TAR) in gRNA and its minus-strand DNA complement, respectively [18,19,40]. ...
... During the course of virus replication, the HIV-1 NC protein interacts with structured RNA elements present at the 5′ end of the viral genome: the TAR stem-loop (SL), which is at the extreme 5′ end of R in gRNA (Fig. 2a) and is involved in the minus-strand transfer step of reverse transcription (reviewed in refs. [18,19,40]); and the Psi region, composed of three SLs, which contributes to the dimerization and packaging of gRNA and has been studied extensively [25,59,60]. The SL structures include SL1, which contains the dimerization initiation site (DIS) and two bulges, SL2 containing the major 5′ splice donor site; and SL3, which is important for packaging viral RNA (Fig. 2b). ...
Article
Full-text available
Background The nucleocapsid (NC) domain of HIV-1 Gag is responsible for specific recognition and packaging of genomic RNA (gRNA) into new viral particles. This occurs through specific interactions between the Gag NC domain and the Psi packaging signal in gRNA. In addition to this critical function, NC proteins are also nucleic acid (NA) chaperone proteins that facilitate NA rearrangements during reverse transcription. Although the interaction with Psi and chaperone activity of HIV-1 NC have been well characterized in vitro, little is known about simian immunodeficiency virus (SIV) NC. Non-human primates are frequently used as a platform to study retroviral infection in vivo; thus, it is important to understand underlying mechanistic differences between HIV-1 and SIV NC. ResultsHere, we characterize SIV NC chaperone activity for the first time. Only modest differences are observed in the ability of SIV NC to facilitate reactions that mimic the minus-strand annealing and transfer steps of reverse transcription relative to HIV-1 NC, with the latter displaying slightly higher strand transfer and annealing rates. Quantitative single molecule DNA stretching studies and dynamic light scattering experiments reveal that these differences are due to significantly increased DNA compaction energy and higher aggregation capability of HIV-1 NC relative to the SIV protein. Using salt-titration binding assays, we find that both proteins are strikingly similar in their ability to specifically interact with HIV-1 Psi RNA. In contrast, they do not demonstrate specific binding to an RNA derived from the putative SIV packaging signal. Conclusions Based on these studies, we conclude that (1) HIV-1 NC is a slightly more efficient NA chaperone protein than SIV NC, (2) mechanistic differences between the NA interactions of highly similar retroviral NC proteins are revealed by quantitative single molecule DNA stretching, and (3) SIV NC demonstrates cross-species recognition of the HIV-1 Psi RNA packaging signal.
... It has been proposed that the primer tRNA Lys3 simultaneously binds the PBS (primer binding sequence) at the 5 end and the S1 sequence (5 -GCCUGGGCGGGACU-3 ) located at the 3 end. This dual interaction induces a circularization of the RNA genome that brings the two R regions into proximity and triggers the minus-strand transfer (78)(79)(80). Interestingly, the S1 sequence is embedded in the U3-G4 studied in this work (Table 1). ...
Article
Full-text available
G-quadruplexes (G4s) are four-stranded nucleic acid structures formed by the stacking of G-tetrads. Here we investigated their formation and function during HIV-1 infection. Using bioinformatics and biophysics analyses we first searched for evolutionary conserved G4-forming sequences in HIV-1 genome. We identified 10 G4s with conservation rates higher than those of HIV-1 regulatory sequences such as RRE and TAR. We then used porphyrin-based G4-binders to probe the formation of the G4s during infection of human cells by native HIV-1. The G4-binders efficiently inhibited HIV-1 infectivity, which is attributed to the formation of G4 structures during HIV-1 replication. Using a qRT-PCR approach, we showed that the formation of viral G4s occurs during the first 2 h post-infection and their stabilization by the G4-binders prevents initiation of reverse transcription. We also used a G4-RNA pull-down approach, based on a G4-specific biotinylated probe, to allow the direct detection and identification of viral G4-RNA in infected cells. Most of the detected G4-RNAs contain crucial regulatory elements such as the PPT and cPPT sequences as well as the U3 region. Hence, these G4s would function in the early stages of infection when the viral RNA genome is being processed for the reverse transcription step.
... Previous work from our group showed that dNTPs increase the integration rate in rCD4 T cells without inducing T-cell activation (11). We also hypothesize that pausing of the reverse transcriptase (RT) on the HIV template contributes to the formation of deleted preintegration sequences, increasing the chance that RT could dissociate and then reassociate in a different region of the template, potentially leading to the formation of deleted viral sequences (33,34). ...
Article
Full-text available
The major implication of our work is that the decay of intact proviruses in vitro is extremely rapid, perhaps as a result of enhanced expression. Gaining a better understanding of why intact proviruses decay faster in vitro might help the field identify strategies to purge the reservoir in vivo . When used wisely, in vitro models are a powerful tool to study the selective pressures shaping the viral landscape. Our finding that massively deleted sequences rarely succeed in integrating has several ramifications. It demonstrates that the total HIV DNA can differ substantially in character from the integrated HIV DNA under certain circumstances. The presence of unintegrated HIV DNA has the potential to obscure selection pressures and confound the interpretation of clinical studies, especially in the case of trials involving treatment interruptions.
... In addition, strand transfer and further elongation of the synthesized DNA can be facilitated by the presence of viral and cellular proteins. Thus, it has been shown that the nucleic acid annealing and helix destabilization activities of the viral nucleocapsid (NC) protein promote strand transfer (11)(12)(13)(14)(15). On the other hand, cytoplasmic lysates of Jurkat cells were found to increase template switching although the specific protein responsible for these effects was not identified (16). ...
Article
Full-text available
During reverse transcription of the HIV-1 genome, two strand-transfer events occur. Both events rely on the RNase H cleavage activity of reverse transcriptases (RTs) and template homology. Using a panel of mutants of HIV-1BH10 (group M/subtype B) and HIV-1ESP49 (group O) RTs and in vitro assays, we demonstrate that there is a strong correlation between RT minus-strand transfer efficiency and template-primer binding affinity. The highest strand transfer efficiencies were obtained with HIV-1ESP49 RT mutants containing the substitutions K358R/A359G/S360A, alone or in combination with V148I or T355A/Q357M. These HIV-1ESP49 RT mutants had been previously engineered to increase their DNA polymerase activity at high temperatures. Now, we found that RTs containing RNase H-inactivating mutations (D443N or E478Q) were devoid of strand transfer activity, while enzymes containing F61A or L92P had very low strand transfer activity. The strand transfer defect produced by L92P was attributed to a loss of template-primer binding affinity, and more specifically, to the higher dissociation rate constants (koff) shown by RTs bearing this substitution. Although L92P also deleteriously affected the RT's nontemplated nucleotide addition activity, neither nontemplated nucleotide addition activity nor the RT's clamp activities contributed to increased template switching when all tested mutant and wild-type RTs were considered. Interestingly, our results also revealed an association between efficient strand transfer and the generation of secondary cleavages in the donor RNA, consistent with the creation of invasion sites. Exposure of the elongated DNA at these sites facilitate acceptor (RNA or DNA) binding and promote template switching.
... Like the forced template switches induced on the RIO templates used here, minus-strand transfer involves homologous donor and acceptor templates on single RNAs. However, accurate minus-strand transfer is guided by factors in addition to primer-template complementarity [44][45][46]. Notably, acceptor template sequences downstream of the growing point for DNA synthesis, which are essential to the "acceptor template invasion" step of retroviral recombination, are completely dispensable for strong stop transfer [37,47]. Thus, rather than study mismatch extension during strong stop transfer, where factors other than primer-template complementarity might mask the properties of interest here, RIO vectors were developed to force template switching at non-strong stop sites, to more accurately recapitulate the recombinogenic template switches that occur about once per kilobase during retroviral DNA synthesis. ...
Article
A key contributor to HIV-1 genetic variation is reverse transcriptase errors. Some mutations result because reverse transcriptase (RT) lacks 3' to 5' proofreading exonuclease and can extend mismatches. However, RT also excises terminal nucleotides to a limited extent, and this activity contributes to AZT resistance. Because HIV-1 mismatch resolution has been studied in vitro but only indirectly during replication, we developed a novel system to study mismatched basepair resolution during HIV-1 replication in cultured cells, using vectors that force template switching at defined locations. These vectors generated mismatched reverse transcription intermediates, with proviral products diagnostic of mismatch resolution mechanisms. Outcomes for wild-type (WT) RT and an AZT-resistant (AZT(R)) RT containing a thymidine analog mutation set -D67N, K70R, D215F, K219Q-were compared. AZT(R) RT did not excise terminal nucleotides more frequently than WT, and for the majority of tested mismatches, both WT and AZT(R) RTs extended mismatches in more than 90% of proviruses. However, striking enzyme-specific differences were observed for one mispair, with WT RT preferentially resolving dC-rC pairs either by excising the mismatched base or switching templates prematurely, while AZT(R) RT primarily misaligned the primer strand, causing deletions via dislocation mutagenesis. Overall, the results confirmed HIV-1 RT's high capacity for mismatch extension during virus replication, and revealed dramatic differences in aberrant intermediate resolution repertoires between WT and AZT(R) RTs on one mismatched replication intermediate. Correlating mismatch extension frequencies observed here with reported viral mutation rates suggests a complex interplay of nucleotide discrimination and mismatch extension drives HIV-1 mutagenesis.
... It is likely that the virus uses the two pathways and perhaps a third involving the polyA and cpolyA hairpins. Indeed, our results are consistent with the 'acceptor invasion' model proposed by Bambara and co-workers (60,61). In this model, RT pausing at the base of the 5' TAR hairpin initiates RNases H cleavages in the 5' polyA hairpin creating a gap for the invasion of the 3' polyA hairpin (acceptor RNA) and interaction with the cpolyA hairpin. ...
Article
Full-text available
An essential step of human immunodeficiency virus type 1 (HIV-1) reverse transcription is the first strand transfer that requires base pairing of the R region at the 3'-end of the genomic RNA with the complementary r region at the 3'-end of minus-strand strong-stop DNA (ssDNA). HIV-1 nucleocapsid protein (NC) facilitates this annealing process. Determination of ssDNA structure is needed to understand the molecular basis of NC-mediated genomic RNA-ssDNA annealing. For this purpose, we investigated ssDNA using structural probes (nucleases and potassium permanganate). This study is the first to determine the secondary structure of the full-length HIV-1 ssDNA in the absence or presence of NC. The probing data and the phylogenetic analysis support the folding of ssDNA into three stem-loop structures and the presence of four high-affinity binding sites for NC. Our results support a model for the NC-mediated annealing process in which the preferential binding of NC to four sites triggers unfolding of the three-dimensional structure of ssDNA, thus facilitating interaction of the r sequence of ssDNA with the R sequence of the genomic RNA. In addition, using gel retardation assays and ssDNA mutants, we show that the NC-mediated annealing process does not rely on a single pathway (zipper intermediate or kissing complex).
... We speculate here the presence of a still unknown mechanism implicated in the generation of the viral diversity described in our results. This mechanism would act concomitantly with the initial steps of reverse transcription, after the first minus strand DNA transfer since the observed changes are strongly concentrated at the 3' end of proviral DNA [42]. The majority of observed chances were nucleotide substitutions. ...
Article
Full-text available
In order to establish new infections HIV-1 particles need to attach to receptors expressed on the cellular surface. HIV-1 particles interact with a cell membrane receptor known as CD4 and subsequently with another cell membrane molecule known as a co-receptor. Two major different co-receptors have been identified: C-C chemokine Receptor type 5 (CCR5) and C-X-C chemokine Receptor type 4 (CXCR4) Previous reports have demonstrated cellular modifications upon HIV-1 binding to its co-receptors including gene expression modulations. Here we investigated the effect of viral binding to either CCR5 or CXCR4 co-receptors on viral diversity after a single round of reverse transcription. CCR5 and CXCR4 pseudotyped viruses were used to infect non-stimulated and stimulated PBMCs and purified CD4 positive cells. We adopted the SOLiD methodology to sequence virtually the entire proviral DNA from all experimental infections. Infections with CCR5 and CXCR4 pseudotyped virus resulted in different patterns of genetic diversification. CCR5 virus infections produced extensive proviral diversity while in CXCR4 infections a more localized substitution process was observed. In addition, we present pioneering results of a recently developed method for the analysis of SOLiD generated sequencing data applicable to the study of viral quasi-species. Our findings demonstrate the feasibility of viral quasi-species evaluation by NGS methodologies. We presented for the first time strong evidence for a host cell driving mechanism acting on the HIV-1 genetic variability under the control of co-receptor stimulation. Additional investigations are needed to further clarify this question, which is relevant to viral diversification process and consequent disease progression.
... NCp7 is a small, basic nucleic acid binding protein containing two zinc-binding domains, i.e., zinc fingers (ZFs), each with the invariant CCHC motif, which are connected by a short basic linker peptide (Darlix et al., 1995(Darlix et al., , 2011Levin et al., 2005Levin et al., , 2010Rein et al., 1998;Thomas and Gorelick, 2008). NCp7 and the NC domain in Gag are essential for multiple events in the virus life cycle including viral RNA dimerization and packaging, virus assembly, reverse transcription, and integration (reviewed in Darlix et al., 2011;Isel et al., 2010;Levin et al., 2005Levin et al., , 2010Lyonnais et al., 2013;Mirambeau et al., 2010;Piekna-Przybylska and Bambara, 2011;Rein et al., 1998;Sleiman et al., 2012;Thomas and Gorelick, 2008). Importantly, NCp7 is a nucleic acid chaperone, i.e., it remodels nucleic acid structures to form the most thermodynamically stable NCp15,NCp9,p6,NCp7,and SP2, are also shown. ...
Article
Human immunodeficiency virus (HIV) stores its genetic information in the form of RNA. This RNA genome is introduced into the target cell during infection. The virus belongs to the family of retroviruses, as it is able to reverse the normal flow of genetic information from DNA to RNA by copying its RNA genome into DNA using the viral enzyme reverse transcriptase (RT). Each viral particle contains two copies of positive-strand RNA genome enclosed by a core composed of 2,000 copies of the capsid (CA) protein. The RNA genome is tightly bound to nucleocapsid proteins (NC) and other enzymes needed for the early steps of viral infection and is protected by a capsid surrounded by a shell composed of matrix (MA) proteins. The shell is located underneath the virion envelope, a plasma membrane of host-cell origin. During infection, the viral envelope fuses with the cell membrane, releasing the capsid into the cytoplasm. Thereafter, the RNA genome undergoes a multistep process of conversion into DNA within the reverse transcription complex (RTC) (Fig. 2.1). After reverse transcription, the newly synthesized DNA is integrated into the host nuclear genome and permanently linked with their target cells. © 2013 Springer Science+Business Media New York. All rights are reserved.
Article
The 1st strand transfer, a crucial step of reverse transcription involving the HIV-1 nucleocapsid protein (NC), relies on base pairing of the r sequence of strong-stop DNA (ssDNA) with the 3' R sequence of viral RNA (3' UTR) which forms the TAR and polyA stem-loops. The r sequence can form the cTAR and cpolyA stem-loops. Therefore, the transfer relies probably on annealing of folded molecules. This process is not well known at the molecular and structural level. The tools of molecular biology and three DNA-targeted probes were used to get insights into the annealing process. Our results were the following: 1) in the absence of NC, the cTAR DNA folds into two distinct conformations in equilibrium; 2) NC slightly shifts the equilibrium toward one conformation and binds tightly the internal loop of the cTAR hairpin; 3) NC is required for the formation of heteroduplex of the full-length ssDNA and 3' UTR; 4) the annealing of ssDNA to 3' UTR can be initiated from different sites in the presence of 0.2 mM MgCl2; 5) the full-length ssDNA folds into two conformations in equilibrium in 0.2 mM MgCl2 but mainly into one conformation in 2 mM MgCl2 ; 6) NC preferentially binds to the single-stranded region between the cTAR and cpolyA hairpins in ssDNA. This binding site probably plays an important role in the annealing of complementary DNA and RNA hairpins. This study helps us to gain insights into the reverse transcription process and the associated genetic recombination.
Article
Full-text available
The Publishers wish to apologize for an incorrect presentation of text accompanying this paper. The peptide structures featured in ‘Table I’ should appear in the text of the ‘Materials and Methods’ under the section entitled ‘Preparation of NCp7 and NCp7 mutants’. The correct version of Table I is reprinted below.
Article
Full-text available
Two obligatory DNA strand transfers take place during reverse transcription of a retroviral RNA genome. The first strand transfer is facilitated by terminal repeat (R) elements in the viral genome. This strand-transfer reaction depends on base pairing between the cDNA of the 5'R and the 3'R. There is accumulating evidence that retroviral R regions contain features other than sequence complementarity that stimulate this critical nucleic acid hybridization step. The R region of the human immunodeficiency virus type 1 (HIV-1) is relatively extended (97 nt) and encodes two well-conserved stem-loop structures, the TAR and poly(A) hairpins. The role of these motifs was studied in an in vitro strand-transfer assay with two separate templates, the 5'R donor and the 3'R acceptor, and mutants thereof. The results indicate that the upper part of the TAR hairpin structure in the 5'R donor is critical for efficient strand transfer. This seems to pose a paradox, as the 5'R template is degraded by RNase H before strand transfer occurs. We propose that it is not the RNA hairpin motif in the 5'R donor, but rather the antisense motif in the ssDNA copy, which can also fold a hairpin structure, that is critical for strand transfer. Mutation of the loop sequence in the TAR hairpin of the donor RNA, which is copied in the loop of the cDNA hairpin, reduces the transfer efficiency more than fivefold. It is proposed that the natural strand-transfer reaction is enhanced by interaction of the anti-TAR ssDNA hairpin with the TAR hairpin in the 3'R acceptor. Base pairing can occur between the complementary loops ("loop-loop kissing"), and strand transfer is completed by the subsequent formation of an extended RNA-cDNA duplex.
Article
In this study we demonstrate that the HIV-1 leader RNA exists in two alternative conformations, a branched structure consisting of several well-known hairpin motifs and a more stable structure that is formed by extensive long-distance base pairing. The latter conformation was first identified as a compactly folded RNA that migrates unusually fast in nondenaturing gets. The minimally required domains for formation of this conformer were determined by mutational analysis. The poly(A) end DIS regions of the leader are the major determinants of this RNA conformation. Further biochemical characterization of this conformer revealed that both hairpins are disrupted to allow extensive long-distance base pairing. As the DIS hairpin is known to be instrumental for formation of the HIV-1 RNA dimer, the interplay between formation of the conformer and dimerization was addressed. Formation of the conformer and the RNA dimer are mutually exclusive. Consequently, the conformer must rearrange into a branched structure that exposes the dimer initiation signal (DIS) hairpin, thus triggering formation of the RNA dimer. This structural rearrangement is facilitated by the viral nucleocapsid protein NC. We propose that this structural polymorphism of the HIV-1. leader RNA acts as a molecular switch in the viral replication cycle.
Article
Reverse transcription of human immunodeficiency virus type-1 (HIV-1) genomic RNA is primed by tRNA3Lys, whose 3′ end 18 nucleotides are complementary to the viral primer binding site (PBS). We used chemical and enzymatic probes to test the conformation of the viral RNA and tRNA|rlmbopopnbop|Lys|clobop|3, in their free form and in the HIV-1 RNA/tRNA|rlmbopopnbop|Lys|clobop|3 binary complex. Extensive reactivity changes were observed in both molecules upon formation of the binary complex. In the viral RNA, reactivity changes occurred up to 69 nucleotides upstream and 72 nucleotides downstream of the PBS. A secondary structure model of the HIV-1 RNA/tRNA3Lyscomplex accounting for all probinag data as been constructed. It reveals an unexpectdly complex aand compact pseudoknot-like structure in which most of the anticodon loop, and 3′ strand of the anticodon stem and the 5′ part of the variable loop of tRNA3Lysinteract with viral sequences 12 to 39 nucleotides upstream of the PBS. The core of the binary complex is a complex junction formed by two single-stranded sequences of tRNA3Lys, an intramolec ular viral helix, an intramolecular tRNA helix, and two intermolecular helices formed by the template/primer interaction. This junction probably highly constrains the tertiary structure of the HIV/tRNA3Lyscomplex. Compared to the structure of the free molcules, only the D arm of tRNA3Lysand a small viral stem-loop downstream of the PBS are unaffectd in the binary complex. Sequence comparison reveals that the main characteristics of the binary complex model are conserved among all HIV-1 isolates.
Article
The human immunodeficiency virus type 1 (HIV-1) long terminal repeat (LTR) represents a model promoter system and the identification and characterisation of cellular proteins that interact with this region has provided a basic understanding about both general eukaryotic and HIV-1 proviral transcriptional regulation. To date a large number of sequence-specific DNA–protein interactions have been described for the HIV-1 LTR. The aim of this report is to provide a comprehensive, updated listing of these HIV-1 LTR interactions. It is intended as a reference point to facilitate on-going studies characterising the identity of cellular proteins interacting with the HIV-1 LTR and the functional role(s) of specific regions of the LTR for HIV-1 replication.
Article
Theprocessofreversetranscription ofretroviral genomesbegins withthesynthesis ofashort DNA molecule near the5'endoftheRNA template. Thismolecule, termedminus-strand strong-stop DNA,isthen translocated tothe3'endoftheviral RNA bymeansofarepeated sequence, theRregion, present atbothends ofthetemplate. Thetranslocation should result inthetransfer ofgenetic information fromthe5'R region to the3'R region. We havegenerated aseries ofmutants ofMoloney murineleukemia virus withalterations in theR regions byinvitro mutagenesis ofa cloned DNA copyoftheviral genome.Thealtered DNAswere introduced into mouse cells bytransfection, andthetranslocation ofthemutations during viral replication was assessed. Somemutations were nottransferred fromthe5'Rregion tothe3'Rregion; these results werenot inaccord withcurrent models for reversetranscription. Theresults canbeexplained ifDNA molecules shorter thanstrong-stop DNA,formed bypremature termination ofsynthesis, aresometimes translocated. A number ofmutants withlarge deletions intheR region weretested andwereable toreplicate withnormalstrong-stop DNA translocation. Thus,short stretches ofhomology can beusedbythevirus tocarryoutstrong-stop translocations. Theretroviruses carry outanunusual reaction soonafter infection: adouble-stranded DNA copyoftheviral genome issynthesized byreverse transcription ofthevirion RNA. Thereaction isknowntobecomplex (Fig. 1).Synthesis begins withtheextension ofatRNAprimer molecule that is basepaired tothetemplate RNA nearits5'end(6)and quickly stops whenthe5'endofthetemplate iscopied. The product ofthis reaction, knownasminus-strand strong-stop DNA,contains somesequences that areunique intheviral RNA,butalsoincludes arepeated sequence, theR region, whichispresent atboththe5'and3'termini oftheviral RNA (3). Subsequent extension ofthis short DNA requires aextraordinary step: thetranslocation or"jump"ofthe strong-stop DNA tothe3'endofthetemplate. Themecha- nismofthetranslocation isnotwell understood, butpresum- ablyinvolves basepairing ofthestrong-stop DNA totheR region atthe3'acceptor endofthetemplate. After thejump hasoccurred, DNA synthesis resumes andresults inthe formation ofanearly full-length minus-strand DNA. Thecompletion oftheprocess ofreverse transcription requires several moresteps. Theviral RNA isdegraded by theRNaseH activity ofthereverse transcriptase enzyme. Next,synthesis ofthesecond DNA strand isbegunwitha plus-strand strong-stop DNA (15). Thismolecule isinitiated atapolypurine stretch inunique viral sequences attheedge oftheU3region atthe3'endofthegenome; DNA synthesis stops after copying partofthetRNAatthe5'endofthe minus-strand strong-stop DNA.A second jumpoccurs, in whichtheplus-strand strong-stop DNA ispaired withse- quences atthe3'endoftheminusstrand toallowits extension. After extension ofbothstrands, theproduct is complete: afully duplex linear DNA containing aduplication ofsequences called longterminal repeats (LTRs). Thelinear DNA isultimately transported tothenucleus andcircular- ized, andoneormoreoftheviral DNAsareintegrated into thehostchromosome.
Article
The 3′ non-coding region (3′NCR) of strains of dengue 1 (DEN 1), DEN 2, DEN 3, and DEN 4 viruses, isolated in different geographical regions, was sequenced and compared to published sequences of the four dengue viruses. A total of 50 DEN 2 strains was compared: 7 West African strains, 3 Indonesian mosquito strains, 1 Indonesian macaque isolate, and 39 human isolates from Southeast Asia, the South Pacific, and the Caribbean and Americas. Nucleotide sequence alignment revealed few deletions and no repeat sequences in the 3′ NCR of DEN 2 viruses and showed that much of the 3′ NCR was well conserved. The strains could be divided into two groups, sylvatic and human/mosquito/macaque, based on nucleotide sequence homology. A hypervariable region was identified immediately following the NS5 stop codon, which involved a 2–10 nucleotide deletion in human, mosquito, and macaque isolates compared with the sylvatic strains. The DEN 2 3′NCR was also compared with 3′NCR sequences from strains of DEN 1, DEN 3, and DEN 4 viruses. DEN 1 was found to have four copies of an eight nucleotide imperfect repeat following the NS5 stop codon, while DEN 4 virus had a deletion of 75 nucleotides in the 3′NCR. We propose that the variation in nucleotide sequence in the 3′NCR may have evolved as a function of DEN virus transmission and replication in different mosquito and non-human primate/human host cycles. The results from this study are consistent with the hypothesis that DEN viruses arose from sylvatic progenitors and evolved into human epidemic strains. However, the data do not support the hypothesis that variation in the 3′NCR correlates with DEN virus pathogenesis.
Article
Retroviral nucleocapsid (NC) protein is an integral part of the virion nucleocapsid where it coats the dimeric RNA genome. Due to its nucleic acid binding and annealing activities, NC protein directs the annealing of the tRNA primer to the primer binding site and greatly facilitates minus strand DNA elongation and transfer while protecting the nucleic acids against nuclease degradation. To understand the role of NCp7 in viral DNA synthesis, we examined the influence of NCp7 on self-primed versus primer-specific reverse transcription. The results show that HIV-1 NCp7 can extensively inhibit self-primed reverse transcription of viral and cellular RNAs while promoting primer-specific synthesis of proviral DNA. The role of NCp7 vis-à-vis the presence of mutations in the viral DNA during minus strand elongation was examined. NCp7 maximized the annealing between a cDNA(−) primer containing one to five consecutive errors and an RNA representing the 3′ end of the genome. The ability of reverse transcriptase (RT) in the presence of NCp7 to subsequently extend the mutated primers depended upon the position of the mismatch within the primer : template complex. When the mutations were at the polymerisation site, primer extension by RT in the presence of NCp7 was very high, about 40% for one mismatch and 3% for five consecutive mismatches. Mutations within the DNA primer or at its 5′ end had little effect on the extension of viral DNA by RT. Taken together these results indicate that NCp7 plays major roles in proviral DNA synthesis within the virion core due to its ability to promote prime-specific proviral DNA synthesis while concurrently inhibiting non-specific reverse transcription of viral and cellular RNAs. Moreover, the observation that NCp7 enhances the incorporation of mutations during minus strand DNA elongation favours the notion that NCp7 is a factor contributing to the high mutation rate of HIV-1.