Abstract
The APOBEC3 family comprises seven cytidine deaminases (APOBEC3A [A3A] to A3H), which are expressed to various degrees in HIV-1 susceptible cells. The HIV-1 Vif protein counteracts APOBEC3 restriction by mediating its degradation by the proteasome. We hypothesized that Vif proteins from various HIV-1 subtypes differ in their abilities to counteract different APOBEC3 proteins. Seventeen Vif alleles from seven HIV-1 subtypes were tested for their abilities to degrade and counteract A3G, A3F, and A3H haplotype II (hapII). We show that most Vif alleles neutralize A3G and A3F efficiently but display differences with respect to the inhibition of A3H hapII. The majority of non-subtype B Vif alleles tested presented some activity against A3H hapII, with two subtype F Vif variants being highly effective in counteracting A3H hapII. The residues required for activity were mapped to two residues in the amino-terminal region of Vif (positions 39F and 48H). Coimmunoprecipitations showed that these two amino acids were necessary for association of Vif with A3H hapII. These findings suggest that the A3H hapII binding site in Vif is distinct from the regions important for A3G and A3F recognition and that it requires specific amino acids at positions 39 and 48. The differential Vif activity spectra, especially against A3H hapII, suggest adaptation to APOBEC3 repertoires representative of different human ancestries. Phenotypic assessment of anti-APOBEC3 activity of Vif variants against several cytidine deaminases will help reveal the requirement for successful replication in vivo and ultimately point to interventions targeting the Vif-APOBEC3 interface.
INTRODUCTION
Apolipoprotein B mRNA-editing, catalytic polypeptide (APOBEC3) cytidine deaminases act as intracellular inhibitors of HIV-1 replication (1, 8). The family comprises seven proteins (APOBEC3A [A3A] to A3H), which differ in catalytic activity and susceptibility to HIV-1 Vif-mediated degradation (1, 8, 38). In the absence of a functional Vif, some APOBEC3 proteins (e.g., A3G, A3F, and A3H haplotype II [hapII]) exert restriction through editing- and nonediting-based mechanisms, thereby preventing efficient viral spread in the next round of infection (1, 8, 13). HIV-1 isolates harbor a Vif allele that degrades A3G and some other antiretroviral APOBEC3 proteins in a ubiquitin/proteasome-dependent manner (1, 8, 38).
A3G and A3F contain two deaminase domains while the A3H protein consists of only one catalytic domain. Moreover, A3H has several haplotypes, some of which are highly active against HIV-1 (haplotypes II, V, and VII) (5, 9, 18, 19, 29, 43, 52). Interestingly, the allele frequency of the active A3H hapII differs considerably among human ethnicities, being high in African and low in European and Asian populations (29, 43). There remains some uncertainty with respect to the Vif neutralization sensitivity of these A3H protein variants (5, 10, 41, 52).
The global HIV-1 epidemic is mostly driven by non-subtype B virus strains from group M (12), but our current knowledge on functional determinants of Vif-APOBEC3 interactions is almost exclusively based on the interactions of HIV-1 subtype B molecular clones (LAI, HXB2, and NL4-3) with the reference A3G. These studies have generated a body of evidence indicating that the amino-terminal region of Vif is important for binding to A3G and A3F proteins (3, 16, 23, 24, 26, 34, 35, 37, 42, 44). Vif residues that are important for selective association with A3G/A3F vary: Vif residues 14 to 17 (DRMR) are important for interaction with A3F, whereas residues 40 to 44 (YRHHY) are required for A3G binding (26, 34, 35, 49). Other studies showed that Vif residues I9, K22, E45, and N48 are necessary for A3G binding, whereas Q12 is needed for A3F binding (37, 44). In addition, tryptophans at positions 11 and 79 and at positions 5, 21, 38, and 89 are important for A3F and A3G binding, respectively (42). The YXXL (X stands for any residue) motif spanning positions 69 to 72 and the WXSLVK motif from positions 21 to 26 are necessary for the degradation of both A3G and A3F (6, 32).
The C-terminal region of Vif contains the SLQYLA SOCS box (residues 145 to 151), which is important for Vif binding to elongin C in the E3 ubiquitin ligase complex (39, 50). This region is highly conserved among HIV-1 subtypes, and mutations within this domain disrupt the interaction between Vif and Elongin C, thereby abrogating anti-APOBEC3 activity (20, 36). This region also contains the zinc binding HCCH motif consisting of residues H108, C114, C133, and H139, which are necessary for Vif binding to Cullin 5, which mediates polyubiquitination of APOBEC3 proteins, followed by their degradation via the 26S proteasome (21, 25, 31, 47, 48). To date, the three-dimensional structure of Vif has not yet been solved by X ray or nuclear magnetic resonance (NMR), but a conformational analysis of HXB2 Vif by hydrogen exchange mass spectrometry suggested that the N-terminal Vif region is more structured than the C-terminal portion of the Vif protein (22). Taken together, our current understanding suggests that the amino-terminal region of HIV-1 Vif allows APOBEC3-specific interactions while the C-terminal region of Vif connects it to the host cell degradation machinery.
Although Vif variants have been characterized genotypically (2, 40, 45, 46), we know very little about the range of phenotypic activity of non-subtype B Vif variants against the different APOBEC3 proteins. Indeed, naturally occurring single residue substitutions may affect the Vif/APOBEC3 phenotype, with full-length subtype B Vif alleles displaying neutralization defects against A3G, A3F, or both deaminases (27, 37). A recent study compared the anti-A3G activity of Vif alleles derived from various subtypes and recombinants (A, B, C, CRF01, and CRF02) and found that subtype C Vif alleles displayed higher activity against A3G (14). The correlates of the increased activity were mapped to several residues in the amino-terminal region of Vif (14), but our understanding of the requirements of diverse Vif proteins to counteract multiple APOBEC3 proteins remains incomplete.
We hypothesized that Vif alleles, especially those obtained from non-B subtype isolates that adapted to replication in individuals of different ancestries, may differ in their relative activities against A3G, A3F, and A3H hapII. In this study, we show that most Vif alleles from different HIV-1 subtypes have activity against A3G and A3F. In contrast, only half of the Vif alleles tested were partially or fully active against A3H hapII. Two distinct residues in Vif (39F and 48H) were required for the efficient neutralization of A3H hapII by Vif, suggesting that the single deaminase A3H may interact with a nonlinear domain in Vif.
MATERIALS AND METHODS
Cloning.
Twelve viral isolates and five molecular clones from different subtypes were obtained through the NIH AIDS Research and Reference Reagent Program (Table 1). Briefly, viral RNA was extracted from the viral stocks using a Qiagen QIAamp Viral RNA Mini Kit, followed by reverse transcription and amplification using primers located outside the Vif coding region. The PCR fragments were cloned into pSC-A (StrataClone PCR cloning vector; Stratagene), and several clones were sequenced and manually aligned to the published sequences. Clones with Vif sequences identical to the published sequences were used as templates for PCR amplification with Vif primers specific for the 5′ and 3′ regions of each variant and containing NotI and EcoRI restriction sites. PCR fragments were digested with NotI/EcoRI, and the coding regions of the subcloned 17 Vif variants were inserted in pCRV1 as described previously (37, 51). Site-directed mutagenesis of subtype F1 and F2 Vifs was performed using overlapping PCR as described previously (37). The mutated Vifs were cloned into pCRV1 vector and confirmed by sequencing. Primer sequences are available upon request.
Table 1.
Description of the Vif alleles used in this study
Vif name | Strain name | Country of origin |
---|---|---|
A1 | 93RW037 | Rwanda |
A2 | 92RW008 | Rwanda |
B-NL4-3 | pNL4-3 | USA |
B-LAI | pLAI.2 | USA |
B1 | 92TH026 | Thailand |
B2 | AD.MDR01 | USA |
B3 | 93TH305 | Thailand |
C1 | 92ZW101 | Zimbabwe |
C2 | 92ZW106 | Zimbabwe |
C3 | pMJ4 | Botswana |
D1 | p94UG114.1.6 | Uganda |
D2 | 94KE102 | Kenya |
AE1 | p90CF402.1.8 | Central African Republic |
F1 | 93BR029 | Brazil |
F2 | 93BR019 | Brazil |
F3 | 93BR020 | Brazil |
G1 | HIV-1 G3 | Nigeria |
Amino-terminally FLAG-tagged A3G, A3F, and A3H hapII in PTR600 expression plasmids have been described previously (10, 28, 30). A carboxyl-terminal triple hemagglutinin (HA) tag was added to A3G and A3H hapII by overlapping PCR, followed by cloning into PTR600. An untagged version of A3H hapII was cloned into PTR600. All DNA preparations were sequenced to confirm the integrity of the APOBEC3 sequences. NCBI reference sequence numbers for the APOBEC3 used are NM 021822.3 (A3G), NM 145298.5 (A3F), and FJ376614.1 (A3H hapII, splice variant SV-183).
Cell culture.
HEK 293T and TZM-bl cells were grown in Dulbecco's modified Eagle's medium (Invitrogen) supplemented with 100 U/ml penicillin/streptomycin and 10% fetal bovine serum (FBS). TZM-bl cells were provided by J. C. Kappes and X. Wu through the AIDS Research and Reference Reagent Program, Division of AIDS, NIAID, National Institutes of Health, NIH Reagent program.
Infectivity assay.
To determine the level of infectivity rescue achieved by each Vif variant, viral stocks were produced in the presence and absence of APOBEC3. Briefly, 3.0 × 105 HEK 293T cells were cotransfected with 500 ng of pNL4-3ΔVif, 5 ng of Vif, and 50 ng of PTR600FLAG-A3G, 50 ng of PTR600FLAG-A3F, or 25 ng of PTR600FLAG-A3H expression plasmids, using 4 μg/ml polyethylenimine (PEI; Polysciences, Inc.). The total amount of DNA transfected was 1,000 ng, with pcDNA3.1 serving as filler. Viral supernatants were collected at 44 to 48 h posttransfection, clarified by centrifugation, and stored at −80°C. Forty-microliter supernatants were used to infect 1 × 104 TZM-bl cells in black 96-well plates. Infections were done in triplicate with viral supernatants from three independent experiments for each APOBEC3 protein tested. Infectivity of the virus particles in the TZM-bl cells was assessed at 44 to 48 h postinfection using a Galacto-Star System for detecting β-galactosidase activity (Applied Biosystems), as described previously (10, 37).
Vif expression.
A total of 3.0 × 105 HEK 293T cells were transfected with 200 ng of Vif expression plasmids, 100 ng of PTR600-green fluorescent protein (GFP), and 700 ng of pcDNA (24-well plate; the total amount of DNA transfected was 1,000 ng with 4 μg/ml PEI). SDS lysis buffer (1% SDS, 50 mM Tris-HCl, pH 8.0, 150 mM NaCl, 5 mM EDTA) was added to the transfected cells 48 h later, and the clarified lysates were analyzed by Western blotting.
Degradation assay.
A total of 3.0 × 105 HEK 293T cells in 24-well plates were cotransfected with 500 ng of pNL4-3ΔVif, 100 ng of PTR600-GFP, 20 ng of pCRV1-Vif, and 100 ng of PTR600FLAG-A3G, 100 ng of PTR600FLAG-A3F, or 50 ng of PTR600FLAG-A3H hapII expression plasmids (total amount of DNA transfected, 1,000 ng; filler DNA, pcDNA3.1).
The degradation of tagged and untagged A3H hapII was compared by cotransfecting HEK 293T cells as described above with the following expression plasmids: PTR600FLAG-A3H hapII (50 ng) and pCRV1-Vif (50 ng) or PTR600-A3H hapII (untagged; 100 ng) with pCRV1-Vif (100 ng). The total amount of DNA transfected was 1,000 ng (filler DNA, pcDNA). Transfections were performed with PEI (4 μg/ml). At 48 h posttransfection cells were lysed (in lysis buffer consisting of 1% SDS, 50 mM Tris-HCl, pH 8.0, 150 mM NaCl, 5 mM EDTA), clarified by centrifugation, and analyzed by Western blotting.
Western blotting.
Transfected HEK 293T cell lysates were separated on 10% bis-Tris 26-well gels (Invitrogen) and transferred by semidry transfer onto polyvinylidene difluoride (PVDF) membranes (Thermo Scientific). Membranes were blocked in 1% (wt/vol) milk solution for 1 h and incubated in primary antibody overnight. HIV-1 Vif was detected with a rabbit anti-Vif primary antibody (NIH catalog number 2221) at 1:1,000 dilution and with goat anti-rabbit secondary antibody (1:5,000; Sigma). FLAG-tagged APOBEC3 proteins were detected with a mouse anti-FLAG primary antibody (1:5,000; Sigma), GFP was detected with rabbit anti-GFP primary antibody (1:2,000; Santa Cruz), and glyceraldehyde-3-phosphate dehydrogenase (GAPDH) was detected with a mouse anti-GAPDH primary antibody (1:2,000; Santa Cruz). Goat anti-rabbit or goat anti-mouse secondary antibody (both, Sigma) was at a 1:5,000 dilution. Membranes were washed in wash buffer (0.1% Tween in phosphate-buffered saline [PBS]) and incubated in secondary antibody for 2 h. Membranes were developed with SuperSignal West Pico or Femto substrate (Thermo Scientific) in an LAS-3000 imaging system (FujiFilm). Only nonsaturated signals were acquired for further analysis with ImageGauge, version 4.0, software.
Vif-A3H coimmunoprecipitation.
A total of 7 × 105 HEK 293T cells were transfected with pCRV1-Vif expression plasmids (200 ng), PTR600HA-A3H hapII (800 ng), PTR600HA-A3G (800 ng), or PTR600-GFP (800 ng) using Fugene 6 according to the manufacturer's instructions. At 24 h after transfection, clasto-lactacystin β-lactone (10 μM; Sigma) was added to the transfection to prevent potential Vif-mediated A3H degradation. At 48 h after transfection, cells were washed with PBS and lysed on ice for 15 min in lysis buffer (1% Triton X-100 in PBS supplemented with EDTA-free protease inhibitor cocktail [Roche]). Proteins were clarified by centrifugation at 4°C (8,200 × g for 10 min at 4°C) and incubated with anti-HA-coated beads (Easyview; Sigma) for 1 h at 4°C. Beads were extensively washed with lysis buffer supplemented with 75 mM NaCl. Bound proteins were eluted by boiling the beads in sample loading buffer. Cell lysates and bound proteins were analyzed by Western blotting.
A3H polyclonal serum.
Two rabbits were immunized three times with the A3H peptide Gln-Phe-Asn-Asn-Lys-Arg-Arg-Leu-Arg-Arg-Pro-Tyr-Tyr-Pro-Arg (A3H residues 12 to 26). Custom antibody production was performed by Pocono Rabbit Farm and Laboratory, Inc. Rabbit polyclonal serum was used at a 1:2,000 dilution to detect transfected untagged and tagged A3H.
Statistical analysis.
GraphPad Prism, version 5, software was used to perform all statistical tests (minimum, maximum, mean, t tests, and Spearman rank correlation). P values are two sided, and values of <0.05 were considered to be significant.
RESULTS
Genotype and expression of selected Vif alleles.
Inspection of >1,500 Vif alleles deposited at the Los Alamos HIV sequence database revealed considerable variation in Vif sequences. Some subtype-specific motifs such as 8L (subtype C), 45DCXH48 (subtype D), and 91QRK93 (subtype AE) were also identified. We selected 12 viral isolates from patients representing seven HIV-1 subtypes and five replication-competent full-length molecular clones from three different subtypes. The subtype-specific motifs listed above were present in the selected Vif alleles. The patients from whom the viruses were originally isolated lived in various countries around the world, like Brazil, Rwanda, Zimbabwe, and Thailand (Table 1). Analysis of the phylogenic relationships of the selected Vif alleles and their corresponding consensus sequences confirmed their subtypes (Fig. 1A).
Fig 1.
A panel of 17 Vif alleles from different subtypes. (A) Phylogenetic relationships of subtype consensus sequences and selected Vif alleles from different subtypes were inferred using the neighbor-joining method. Bootstrap values of 70% or higher are shown next to the branches (1,000 replicates). Distances were computed using the maximum composite likelihood method. All positions containing gaps and missing data were eliminated from the data set (complete deletion option). There were a total of 568 positions in the final data set. Phylogenetic analyses were conducted in MEGA4. (B) Vif variants from different subtypes are expressed to comparable levels. pCRV1 Vif expression plasmids were transfected with GFP expression plasmid into HEK 293T cells. Lysates were separated by SDS-PAGE at 48 h posttransfection, transferred to PVDF membranes, and probed with antibodies against Vif, GFP, and GAPDH. (C) Relative Vif expression for each Vif variant is shown. Nonsaturated signals were acquired, background was subtracted, and values were normalized to GFP expression levels. NL4-3 Vif expression was set to 1. The average and standard deviations of three independent experiments are shown. Error bars represent standard deviations. α denotes anti.
At the protein sequence level, the 17 Vif alleles displayed, on average, 15% diversity relative to HIV subtype B molecular clone LAI (maximum, 19%; minimum, 7%). The greatest number of changes were observed for Vif G1 (37/192 residues), while Vif B3 (29/192 residues) was most similar to LAI Vif. Inspection of the N-terminal region revealed polymorphic positions both within and outside the domains known to be relevant for APOBEC3-specific interactions while the HCCH Cullin 5-binding motif and the SLQYLA Elongin C-binding motif were conserved in all 17 Vif alleles (see Fig. S1 in the supplemental material for Vif sequence alignments).
The majority of Vif variants were similarly expressed upon transfection in 293T cells (Fig. 1B and C), with the exception of the subtype C Vif C1, which consistently displayed a 2-fold decrease in signal intensity (Fig. 1C). This difference may reflect lower Vif protein expression or reduced antibody recognition by the polyclonal anti-Vif rabbit serum. All 17 Vif variants encode 192 residues, but some Vif bands migrated differently during gel electrophoresis, likely reflecting variations in protein charges (Fig. 1B, compare Vif F1 to Vif F3 or Vif G1).
Anti-APOBEC3 phenotypes of Vif alleles.
The different Vif proteins were analyzed for rescue of infectivity of NL4-3 with a deletion of Vif (pNL4-3ΔVif) in the presence of A3G, A3F, and A3H hapII. In addition, APOBEC3 degradation by the Vif panel was assessed by cotransfecting the Vifs and APOBEC3, followed by Western blot analysis. Titrations of Vif expression plasmids (5 ng, 100 ng, and 500 ng) and APOBEC3 plasmids (25 ng, 50 ng, and 100 ng) in the presence of NL4-3ΔVif were performed to select assay conditions and plasmid ratios that allowed optimal discrimination of both an increase and a decrease in rescue of infectivity (data not shown). Five nanograms of Vif and 50 ng of A3G or 50 ng of A3F were used in the experiments comparing the anti-A3G/A3F activities of different Vif variants. In addition, similar experiments were performed with Vif-deleted HIV-1 molecular clones pLAI, pMJ4 (subtype C, Vif C3), and p94UG114 (subtype D, Vif D1) to exclude potential isolate and subtype incompatibilities (data not shown). Of note, Capsid p24 production upon transfection was comparable for all viral stocks (data not shown).
Most HIV-1 Vif alleles counteract A3G and A3F restriction.
The infectivity of viral stocks was quantified 48 h posttransfection using TZM-bl reporter cells as described previously (10, 28, 37). An empty pCRV1 plasmid and the NL4-3 Vif mutant C133S, which abrogates cullin 5 binding (C133S), were used as negative controls. Infectivity of NL4-3ΔVif produced in the absence of both APOBEC3 and Vif was set to 100% relative infectivity (Fig. 2A and C; see also Fig. 3A). We chose the widely used molecular clone LAI Vif variant as a reference for comparison to other Vifs (Fig. 1A), and the differences relative to LAI Vif were calculated using a t test (Prism GraphPad).
Fig 2.
Activity of Vif alleles against A3G and A3F. (A) Anti-A3G activities of the different Vif variants. Infectivity of pNL4-3ΔVif cotransfected with Vif alleles and A3G was assessed at 48 h posttransfection on TZM-bl reporter cells. Data shown are representative of three independent experiments. Error bars represent standard deviations. Unpaired t tests were computed to determine whether differences between LAI Vif infectivity and a given Vif variant reached significance (*, P = 0.05; **, P < 0.01; ***, P < 0.001 [Prism software]). (B) Degradation of A3G by the different Vif alleles. Transfected lysates were separated by SDS-PAGE at 48 h posttransfection, transferred to PVDF membranes, and probed with antibodies against FLAG, Vif, and GFP. One representative Western blot is shown. (C) Anti-A3F activities of the different Vif variants. Infectivity of pNL4-3ΔVif cotransfected with Vif alleles and A3G was assessed at 48 h posttransfection on TZM-bl reporter cells. Data shown are representative of three independent experiments. Error bars represent standard deviations. Unpaired t tests were computed to determine whether differences between LAI Vif infectivity and a given Vif variant reached significance (*, P = 0.05; **, P < 0.01; ***, P < 0.001 [Prism software]). (D) Degradation of A3F by the different Vif alleles. Transfected lysates were separated by SDS-PAGE at 48 h posttransfection, transferred to PVDF membranes, and probed with antibodies against FLAG, Vif, and GFP. Representative Western blots are shown.
Fig 3.
Activity of Vif alleles against A3H hapII. (A) Anti-A3H hapII activities of the different Vif variants. Infectivity of pNL4-3ΔVif cotransfected with the Vif alleles and A3H hapII was assessed at 48 h posttransfection on TZM-bl reporter cells. Data shown are representative of three independent experiments. Error bars represent standard deviations. Unpaired t tests were computed to determine whether differences between LAI Vif infectivity and a given Vif variant reached significance (*, P = 0.05; **, P < 0.01; ***, P < 0.001 [Prism software]). No Vif/No A3H, 100% relative infectivity. (B) Degradation of A3H hapII by the different Vif alleles. Transfected lysates were separated by SDS-PAGE at 48 h posttransfection, transferred to PVDF membranes, and probed with antibodies against FLAG, Vif, and GFP. Representative Western blots of one out of three experiments are shown. No Vif/A3H, 100% A3H remaining. (C) A3H hapII degradation efficiency of different Vif variants. A3H hapII signals were quantified and the no-Vif control is set at 100%. Error bars represent standard deviations of three independent A3H hapII degradation assays. (D) Correlation between Vif-mediated rescue of viral infectivity in the presence of A3H hapII and Vif-mediated degradation. Gray symbols identify the controls (as described for A and B). (E) The pattern of A3H hapII-mediated degradation by Vif is independent of the FLAG tag. Selected active and nonactive Vif variants were transfected with FLAG-tagged and untagged A3H hapII expression plasmids. Transfected lysates were separated by SDS-PAGE at 48 h posttransfection and transferred to PVDF membranes. Proteins were detected with polyclonal rabbit serum directed against A3H or Vif. GFP was detected with a monoclonal antibody.
In the absence of Vif or in the presence of the Vif C133S mutant, A3G strongly reduced infectivity to 1%. The reference LAI Vif rescued infectivity from A3G restriction to 60% ± 5%, and the majority of Vif alleles tested were similarly or more active than LAI Vif (Fig. 2A).
Four Vif variants (NL4-3, B2, C1, and G1) displayed low to moderate activities against A3G (relative infectivity, 35% ± 4.5%, 36% ± 2.1%, 10.3% ± 3.5%, and 35% ± 2.5%, respectively), and only one Vif allele (B3) was fully defective against A3G (relative infectivity, 2.7% ± 1.5%). Interestingly, Vif C1, C2, and C3 were expressed to lower levels than the other Vifs, and correcting infectivity data with the Vif expression levels shown in Fig. 2A and B demonstrated that all subtype C Vifs were relatively more active against A3G.
The ability of the Vifs to degrade the APOBEC3s was determined by transfecting HEK 293T cells with 20 ng of the different Vif expression plasmids and 100 ng of A3G (Fig. 2B) or A3F (Fig. 2D). The rescue of infectivity correlated with the ability to degrade A3G in the producer cell, with the exception of Vif C1, which degraded A3G efficiently but failed to rescue viral infectivity to the same degree.
A3F is less potent than A3G in restricting NL4-3ΔVif complemented with an empty control plasmid or with the mutant Vif C133S (no-Vif and Vif C133S, 10% compared to 1% with A3G). LAI Vif rescued infectivity in the presence of A3F to 64% ± 5%, with all Vif alleles displaying activity against A3F. Vif variants AE1, F2, and F3 were more active than LAI Vif (relative infectivity, 73% ± 8.1%, 85% ± 8.5%, and 71% ± 6.6%, respectively). Vif C1, which displayed low activity against A3G, showed average activity against A3F (relative infectivity, 50% ± 2.0%). Interestingly, Vif B3, which was not active against A3G, displayed some activity against A3F (relative infectivity, 26% ± 3.1%) although its ability to degrade A3F was modest (Fig. 2D). Overall, most Vifs were able to degrade and counteract A3G and A3F restriction.
Most non-subtype B HIV-1 Vif alleles counteract A3H hapII restriction.
More than eight naturally occurring A3H haplotypes have been described, with A3H hapII having potent anti-HIV activity (10, 29, 43). There remains some uncertainty about the extent to which A3H is partially or completely resistant to Vif neutralization (5, 10, 29, 41, 43, 52). We therefore tested whether the different Vif alleles have activity against A3H hapII (Fig. 3A). Based on initial optimization experiments (data not shown), we used less APOBEC3 expression plasmid for the experiments with A3H hapII (25 ng of A3H hapII compared to 50 ng of A3G or A3F). Under these experimental conditions, A3H hapII reduces infectivity to 10% ± 1.0% in the absence of Vif or with the mutant Vif C133S.
Interestingly, 9 out of 17 Vif variants were able to rescue infectivity in the presence of A3H hapII. Of the active Vifs, Vif F1 and F3 were, by far, the most efficient at counteracting A3H hapII (F1, 80% ± 7.5%; F3, 77% ± 8.0%) (Fig. 3A), followed by LAI, two subtype C Vif variants (C1 and C3), and one subtype A Vif variant (A1). Of note, Vif C1, which had low activity against A3G and modest activity against A3F, was active against A3H hapII (35% ± 2.0%). Three out of four subtype B variants failed to counteract A3H hapII restriction, with LAI Vif being the sole exception. LAI Vif rescued viral infectivity in the presence of A3H hapII to 43% ± 6.1%, whereas NL4-3 Vif was defective in rescuing A3H-mediated restriction (11% ± 2.6%). This observation is in good agreement with reports showing that Vif from LAI but not NL4-3 is active against A3H hapII (18, 19).
The ability of Vif variants to degrade the A3H hapII was determined by transfecting HEK 293T cells with 20 ng of the different Vif expression plasmids and 50 ng of A3H (Fig. 3B). Overall, A3H hapII was more resistant to degradation than A3G and A3F (Fig. 3B, compare the no-Vif and Vif C133S control levels to the level of A3H in the presence of other Vifs). The Vif variants LAI, F1, and F2 degraded A3H hapII efficiently (Fig. 3C, e.g., only 20% of the “no Vif” A3H hapII control remaining), whereas the other Vif variants were partially or completely defective with respect to degradation in the producer cell (Fig. 3B and C). The degree of Vif-mediated rescue of viral infectivity in the presence of A3H hapII correlated well with the efficiency to degrade A3H hapII (Fig. 3D).
Adding epitope tags to A3H hapII has been reported to influence Vif-mediated A3H hapII degradation (18). To test whether the N-terminal FLAG tag to A3H hapII affects its susceptibility to Vif, we analyzed 10 Vif alleles with N-terminal FLAG tags and untagged A3H hapII for degradation. We transfected the selected Vif variants and the two A3H expression plasmids into HEK 293T cells and used polyclonal rabbit serum raised against an A3H peptide to detect both A3H hapII variants in the cell lysates by Western blotting (Fig. 3E). The A3H hapII degradation patterns were similar for the N-terminally tagged and untagged A3H hapII proteins (Fig. 3E), indicating that the N-terminal FLAG tag did not affect its sensitivity to Vif degradation.
Taken together, in contrast to the efficient activity of most of the Vifs against A3G and A3F, Vifs from different subtypes displayed considerable differences in counteracting A3H hapII restriction.
The spectrum of anti-APOBEC3 activities varies between Vif variants.
Since multiple APOBEC3 proteins are expressed in HIV-1 susceptible cells, we were interested in a comprehensive assessment of Vif-mediated anti-APOBEC3 activities. Figure 4 provides a heat map representation of the anti-APOBEC3 spectrum of each Vif allele (complete rescue indicated in green and maximum restriction shown in red). Vif variants F1 and F3 showed high activity against all three deaminases tested (broad-spectrum anti-APOBEC3 activity), while Vif B3 displayed low activity against all three APOBEC3s. Interestingly, we also noted A3-specific Vif defects in which one (Vif C1 and Vif F2) or two (Vif C2) specific APOBEC3s were not neutralized. Overall, the patterns of rescue from A3G and A3F were similar among Vifs but did not correlate with the activities of these Vifs against A3H hapII. This indicates that residues in Vif that interact with A3G and A3F are likely different from those that interact with A3H hapII.
Fig 4.
Comparison of the anti-APOBEC3 activities of different Vif alleles. A heat map representation illustrates the spectrum of infectivity obtained with each Vif variant in the presence of A3G, A3F, and A3H hapII. Normalized infectivity values are plotted. The controls for each deaminase are shown on the right side of the panel: the level of suppression achieved (%) in the absence of Vif is depicted (red), and the infectivity (%) in the absence of APOBEC3 is set to 100 (green). Relative infectivity values are the average of three independent experiments shown in Fig. 2 and 3. Max denotes maximum.
Identification of the Vif residues required for A3H hapII degradation.
Subtype F Vif variants were efficient in counteracting A3G and A3F but differed in their abilities to block A3H hapII (Fig. 4). This apparent difference between F1 and F2 with respect to A3H hapII was used to accurately determine the Vif requirements for interacting with A3H hapII. We focused on the N-terminal part of Vif because this region has been previously shown to interact with A3G and A3F. Since Vif F1 and Vif F2 differ at only five residues in the N-terminal region (positions 39, 48, 61, 62, and 63) (Fig. 5A), we generated a series of mutants in which residue 39, residue 48, or residues 61 to 63 were exchanged between the two Vifs (Fig. 5A). We tested the infectivity of NL4-3ΔVif with these Vif mutants in the presence and absence of A3H hapII (Fig. 5B). Infectivity data showed that two residues, 39F and 48H, are responsible for the difference between Vif F1 and Vif F2 and that both are required for efficient anti-A3H hapII activity. Replacing either the phenylalanine (F) at position 39 with a serine (39S) or the histidine (H) at position 48 with an asparagine (48N) rendered Vif F1 inactive against A3H hapII, while the introduction of both residues (F39 and H48) was required to render Vif F2 active against A3H hapII (Fig. 5B). The ability of these mutants to degrade A3H hapII was also assessed (Fig. 5C) and showed that the ability to rescue infectivity correlated with A3H hapII degradation (Fig. 5B and C). Interestingly, mutations at these two positions affected only the activity against A3H hapII, whereas the ability to counteract A3G and A3F was unaffected (data not shown).
Fig 5.
Identification of the residues important for activity of subtype F Vif variants against A3H hapII. (A) A schematic of the Vif N-terminal mutants generated to investigate activity against A3H hapII. Vif F1 and Vif F2 differ at positions 39, 48, and 61 to 63. (B) Infectivities of pNL4-3ΔVif cotransfected with Vif F1, Vif F2, and Vif mutants M1 to M8 and A3H hapII were assessed at 48 h posttransfection on TZM-bl reporter cells. Data shown are representative of three independent experiments. Error bars represent standard deviations of triplicate experiments. (C) Western blot depicting A3H hapII degradation in the presence of Vif F1, Vif F2, and Vif mutants M1 to M8. Transfected lysates were separated by SDS-PAGE at 48 h posttransfection, transferred to PVDF membranes, and probed with antibodies against FLAG, Vif, and GFP. Data shown are representative of three independent experiments. (D) Coimmunoprecipitation (Co-IP) of subtype F Vif variants with A3H hapII and A3G. Subtype F Vif F1, Vif F2, or empty control plasmid was cotransfected with HA-tagged A3H hapII, HA-tagged A3G, or control plasmid. APOBEC3 degradation was prevented by the addition of the proteasome inhibitor clasto-lactacystin β. Cleared lysates were incubated with anti-HA-coated beads and washed vigorously. Proteins were analyzed by Western blotting. (E) Coimmunoprecipitation of subtype F Vif variants with A3H hapII. Subtype F Vif F1, Vif F2, and Vif mutants M4 and M3 were cotransfected with HA-tagged A3H hapII. A3H hapII degradation was prevented by the addition of the proteasome inhibitor clasto-lactacystin β. Cleared lysates were incubated with anti-HA-coated beads and washed vigorously. Proteins were analyzed by Western blotting.
Before Vif can mediate the degradation of APOBEC3 by the proteasome, both proteins must first physically interact. We speculated, therefore, that the Vif proteins that failed to counteract A3H hapII might be inefficiently interacting with A3H hapII. We performed coimmunoprecipitation experiments by transfecting Vif F1 and Vif F2 with A3H hapII or A3G in HEK 293T cells. The proteasome inhibitor clasto-lactacystin β-lactone was added to the medium 24 h before lysis to prevent APOBEC3 degradation. Vif F1 was more efficient in precipitating A3H hapII than Vif F2, whereas both Vif F1 and F2 associated efficiently with A3G (Fig. 5D). In addition, we established that the two residues were necessary and sufficient to determine the ability of Vif F variants to associate with A3H hapII by cotransfected Vifs F1 and F2 as well as mutant Vifs M4 and M3 with A3H hapII in HEK 293T cells. Vif F1 and mutant M3 (with residues 39F and 48H) immunoprecipitated with A3H hapII, while defective Vifs F2 and mutant M4 (with 39S and 48N) both failed to associate with A3H hapII. This indicates that a phenylalanine at position 39 and a histidine at position 48 determine the association between subtype F Vifs and A3H hapII (Fig. 5E). These results correlated well with the observed infectivity and degradation phenotypes (Fig. 5B and C).
Alanine scanning of residues 38 to 49 in the amino-terminal region of subtype F Vif.
In addition to the polymorphic nature of the amino acids at position 39 and 48, changes within the region between these two residues may affect the interaction between Vif and APOBEC3. Indeed, several motifs in this region have been described to affect subtype B Vif function against A3G (e.g., 40YRHHY44, 45E, and 48N) (34, 37). We individually replaced all amino acids between positions 38 and 49 in Vif F1 with alanines and analyzed their effects on rescue of infectivity in the presence of A3G, A3F, and A3H hapII. The alanine scanning of this Vif F1 region showed that distinct positions had differential effects on activity against A3G, A3F, and A3H hapII (Fig. 6). Alanines at position 38, 48, and 49 rendered Vif F1 defective against all three APOBEC3s (Fig. 6A, B, C, and E) but did not affect Vif expression (Fig. 6D). None of the other mutations reduced neutralization of A3F, while other mutations showed APOBEC3-specific effects: 40A completely abrogated anti-A3G and anti-A3H hapII activity, and several substitutions in this region resulted in a specific loss of function against A3H hapII (41A, 40A, and 45A) (Fig. 6C). Position 39, one of the two positions that determined the difference between Vif F1 and F2 with regard to A3H degradation, was more tolerant to being mutated to an alanine (from an aromatic phenylalanine) than to the nucleophilic serine. Western blot analysis showed that the efficiency in rescue of infectivity correlates with the level of APOBEC3 degradation by the respective Vifs. Taking these results together, alanine scanning mutagenesis of the region between residues 39 and 48 in subtype F Vif suggests that the residues necessary for A3H hapII recognition are distinct from those required for Vif-A3G or Vif-A3F interactions.
Fig 6.
Characterization of the Vif/APOBEC3 interface that allows degradation of A3H hapII in the Vif F1 context. The anti-APOBEC3 activities of the 12 different alanine mutants in Vif F1 (alanine scanning mutagenesis of the region between residues 38 and 49) are shown. Infectivity of pNL4-3ΔVif cotransfected with the Vif mutants and A3G (A), A3F (B), and A3H hapII (C) was assessed 48 h posttransfection on TZM-bl reporter cells. Data shown are representative of three independent experiments. Error bars represent standard deviations. Western blots showing A3G, A3F, and A3H hapII degradation by the mutant F1 Vifs are directly under the respective infectivity values. (D) Transfected Vif F1 mutants are expressed to comparable levels. Representative Western blots are shown for Vif expression and the GFP transfection control. (E) Heat map summarizing the anti-A3G, -A3F, and -A3H hapII phenotypes associated with each Vif F1 alanine mutant.
Next, we analyzed the sequences between positions 38 and 49 of the initial subtype set of Vifs to understand why only some Vifs counteract A3H hapII. Figure 7 shows an alignment of the protein region of interest (positions 38 to 49) of the 17 Vif proteins tested and their respective activities against A3H hapII. Nearly all Vifs that efficiently counteracted A3H hapII carried the 39F and 48H combination, whereas most of the inactive Vifs carried alternative pairs. Only a few exceptions existed: Vif B2 contains the FH pair but lacks anti-A3H activity, which could be caused by the unique glycine (G) at position 41 (41G, all other active Vifs carry 41R). This possibility is further supported by the alanine scanning results, which showed the importance of position 41 for anti-A3H function (Fig. 6C). Vif D2, which carries an LH pair instead of the FH pair, displayed low activity against A3H hapII. The leucine (L) at position 39 shares the same hydrophobic and nonpolar properties as the phenylalanine (F), indicating that some variation at position 39 is tolerated with amino acids of similar properties, like 39A (Fig. 6C).
Fig 7.
Correlation between Vif genotype and activity against A3H hapII. The amino acid alignment of the 17 Vifs shows the region between residues 38 to 49 in the amino-terminal portion of Vif. Vif F1 serves as reference in this alignment. The Vifs are ranked in descending order of anti-A3H hapII activity (Fig. 4). Nine Vif variants (F1 to G1) rescued viral infectivity in the presence of A3H hapII to 20% or more. These infectivity values were significantly higher (P < 0.05, t test) from the no-Vif or the Vif C133S controls. The level of rescue is indicated as follows: +++, high, with infectivity >50% of the no-A3H control; ++, moderate, with 50 to 30% of the no-A3H control; and +, low, with 20 to 30% of the no-A3H control. Infectivity values of ≤20% of the no-A3H control were defined as lack of neutralization activity (—). The amino acid pairs at locations 39 and 48 are grouped as FH, ZH, and XN. Z denotes any residue except F, and X denotes any residue.
In summary, a 39F-48H or a 39L-48H pair is required for activity against A3H hapII but not for neutralization of A3G and A3F. In addition, amino acid changes at critical residues between these positions (e.g., 40, 41, and 45) may abrogate anti-A3H hapII function in the presence of the FH/LH pairs.
DISCUSSION
At least five cytidine deaminases of the APOBEC3 family restrict retroviruses if left unimpeded by Vif (1, 4, 8, 11, 38). Our knowledge of the anti-APOBEC3 spectra of diverse non-B subtype Vif variants is limited to A3G (14).
In this study, we assessed whether Vifs from different HIV-1 subtype strains are able to counteract specific APOBEC3 proteins (A3G, A3F, and A3H hapII) with comparable efficiencies. We found that several non-B subtype Vif alleles were more efficient in neutralizing A3G, A3F, and A3H hapII than the commonly used subtype B Vifs. Indeed, most Vif alleles tested in this study against A3G and A3F were at least as active as the Vif obtained from the subtype B LAI molecular clone (Fig. 2 and 4). Moreover, the majority of non-B subtypes were able to neutralize A3H hapII (Fig. 3 and 7). The subtype F Vif variants F1 and F3 were especially efficient in neutralizing A3H hapII (>70% of the no-A3H control), while a second group of Vifs (Vif A1, C1, C3, and LAI) displayed moderate activity against A3H hapII (e.g., >30 to 50% infectivity of the no-A3H control). Two residues, 39F and 48H, were required for this activity (Fig. 5). Alanine scanning mutagenesis revealed that changes to other residues between positions 39 and 48 may affect the anti-A3H hapII function even in the presence of the required 39F-48H pair (Fig. 6). The observed variation in anti-A3H hapII activity of the Vifs tested in this study, however, was largely explained by the presence/absence of this specific amino acid pair (Fig. 7). For example, the difference in anti-A3H activities between the LAI and NL4-3 Vifs is likely due to the substitution at position 48, but the additional differences between NL4-3 and LAI Vifs at positions 47 and 50 may also affect anti-A3H activity (Fig. 7). A recent analysis of HIV-1 Vif sequence variation in an Argentinean pediatric cohort (7) showed that approximately 70% of subtype F strains encoded FH at positions 39/48, while 30% carried ZH or XN combinations (where Z indicates any amino acid other than F, and X denotes any residue), suggesting that the majority of subtype F strains likely counteract A3H hapII.
Our findings suggest that HIV-1 Vif variants display a spectrum of neutralization phenotypes, which cannot be easily predicted from the Vif genotype. Some Vifs failed to neutralize A3G but rescued infectivity in the presence of A3F or A3H hap II (e.g., Vif C1), while others displayed activity against A3G and A3F but not A3H hapII. We propose that the sum of these activities likely determines the degree to which APOBEC3 restriction is overcome in the context of given individual's APOBEC3 composition. More comprehensive data sets relating Vif genotype to Vif phenotype are needed to develop algorithms that allow prediction of anti-APOBEC3 activities.
The structure-function experiments performed in this study imply that the Vif interface recognizing A3H hapII is distinct from the A3G or A3F interface. Multiple distant residues are involved, suggesting a nonlinear recognition domain. We speculate that the Vif alleles lacking activity against A3H hapII were originally obtained from individuals who did not express active A3H variants and, thus, did not require A3H neutralization. Interestingly, the anti-APOBEC3 activity of LAI Vif was more potent as well as broader than that of NL4-3 Vif; the LAI Vif not only neutralized A3G and A3F with higher efficiency but also was the only subtype B Vif that displayed activity against A3H hapII. Endogenous A3H expression levels are believed to be modest, but the differential neutralization spectra of several Vif variants of different subtypes support the role of A3H as a restriction factor in vivo.
Some limitations of our study need to be mentioned. First, we cannot draw generalized conclusions about subtype-specific activities since the sample sizes are not sufficiently representative. Moreover, the Vif alleles used in this study were cloned from viral isolates obtained by coculture. Future studies on the anti-APOBEC3 function should use Vif variants directly obtained from the plasma of patients in order to obtain information on the Vif activity spectrum without bias introduced by passaging the viral stock in cells from a different donor. Moreover, experimental approaches using overexpression assays are frequently used to probe for HIV-APOBEC3 interactions, but specific assays using primary T lymphocytes encoding specific APOBEC3 haplotypes are needed to further validate our observations at more physiological APOBEC3 expression levels.
The APOBEC3 locus is polymorphic, with several haplotypes for each of these deaminases (1). Proviruses with footprints of past cytidine deamination are found in many infected patients independently of the HIV-1 subtype (1, 15, 17, 33), suggesting that variation in the spectra of the anti-APOBEC3 activities of HIV-1 Vif and/or antiretroviral activity of Vif-resistant APOBEC3 proteins is important in vivo. HIV-1-infected patients with certain A3H haplotypes displayed fewer G-to-A mutations and lower viral loads (9), and it is attractive to speculate that Vif variants that counteract multiple APOBEC3 proteins decrease the complexity of the viral quasi-species as well as the pathogenicity of HIV-1 in a given infected individual. A previous study found that Vif variants from a patient harboring the active A3H hapII were active against this variant while Vif variants from patients with A3H hapI failed to degrade A3H hapII (19). This observation, together with our findings, suggests a more complex picture in which Vif may adapt to the APOBEC3 haplotype repertoire of the infected host. Future studies looking for clinical associations between APOBEC3 and HIV/AIDS disease progression will have to focus on multiple APOBEC3 deaminases in conjunction with the Vif phenotype.
Supplementary Material
ACKNOWLEDGMENTS
We thank all the members of the Simon laboratory for helpful discussions.
This research was supported by National Institutes of Health grants AI089246, AI90935, and AI064001.
Footnotes
Published ahead of print 19 October 2011
Supplemental material for this article may be found at http://jvi.asm.org/.
REFERENCES
- 1. Albin JS, Harris RS. 2010. Interactions of host APOBEC3 restriction factors with HIV-1 in vivo: implications for therapeutics. Expert Rev. Mol. Med. 12:e4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2. Bell CM, et al. 2007. Molecular characterization of the HIV type 1 subtype C accessory genes vif, vpr, and vpu. AIDS Res. Hum. Retroviruses 23:322–330 [DOI] [PubMed] [Google Scholar]
- 3. Conticello SG, Harris RS, Neuberger MS. 2003. The Vif protein of HIV triggers degradation of the human antiretroviral DNA deaminase APOBEC3G. Curr. Biol. 13:2009–2013 [DOI] [PubMed] [Google Scholar]
- 4. Cullen BR. 2006. Role and mechanism of action of the APOBEC3 family of antiretroviral resistance factors. J. Virol. 80:1067–1076 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5. Dang Y, et al. 2008. Human cytidine deaminase APOBEC3H restricts HIV-1 replication. J. Biol. Chem. 283:11606–11614 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6. Dang Y, Wang X, Zhou T, York IA, Zheng YH. 2009. Identification of a novel WxSLVK motif in the N terminus of human immunodeficiency virus and simian immunodeficiency virus Vif that is critical for APOBEC3G and APOBEC3F neutralization. J. Virol. 83:8544–8552 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7. De Maio FA, et al. 2011. Effect of HIV-1 Vif variability on progression to pediatric AIDS and its association with APOBEC3G and CUL5 polymorphisms. Infect. Genet. Evol. 11:1256–1262 [DOI] [PubMed] [Google Scholar]
- 8. Goila-Gaur R, Strebel K. 2008. HIV-1 Vif, APOBEC, and intrinsic immunity. Retrovirology 5:51. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9. Gourraud PA, et al. 2011. APOBEC3H haplotypes and HIV-1 pro-viral vif DNA sequence diversity in early untreated human immunodeficiency virus-1 infection. Hum. Immunol. 72:207–212 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10. Harari A, Ooms M, Mulder LC, Simon V. 2009. Polymorphisms and splice variants influence the antiretroviral activity of human APOBEC3H. J. Virol. 83:295–303 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11. Harris RS, Liddament MT. 2004. Retroviral restriction by APOBEC proteins. Nat. Rev. Immunol. 4:868–877 [DOI] [PubMed] [Google Scholar]
- 12. Hemelaar J, Gouws E, Ghys PD, Osmanov S. 2011. Global trends in molecular epidemiology of HIV-1 during 2000–2007. AIDS 25:679–689 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13. Hultquist JF, et al. 2011. Human and rhesus APOBEC3D, APOBEC3F, APOBEC3G, and APOBEC3H demonstrate a conserved capacity to restrict Vif-deficient HIV-1. J. Virol. 85:11220–11234 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14. Iwabu Y, et al. 2010. Differential anti-APOBEC3G activity of HIV-1 Vif proteins derived from different subtypes. J. Biol. Chem. 285:35350–35358 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15. Janini M, Rogers M, Birx DR, McCutchan FE. 2001. Human immunodeficiency virus type 1 DNA sequences genetically damaged by hypermutation are often abundant in patient peripheral blood mononuclear cells and may be generated during near-simultaneous infection and activation of CD4[sup]+ T cells. J. Virol. 75:7973–7986 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16. Kao S, et al. 2003. The human immunodeficiency virus type 1 Vif protein reduces intracellular expression and inhibits packaging of APOBEC3G (CEM15), a cellular inhibitor of virus infectivity. J. Virol. 77:11398–11407 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17. Kijak GH, et al. 2008. Variable contexts and levels of hypermutation in HIV-1 proviral genomes recovered from primary peripheral blood mononuclear cells. Virology 376:101–111 [DOI] [PubMed] [Google Scholar]
- 18. Larue RS, Lengyel J, Jonsson SR, Andresdottir V, Harris RS. 2010. Lentiviral Vif degrades the APOBEC3Z3/APOBEC3H protein of its mammalian host and is capable of cross-species activity. J. Virol. 84:8193–8201 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19. LI MM, Wu LI, Emerman M. 2010. The range of human APOBEC3H sensitivity to lentiviral Vif proteins. J. Virol. 84:88–95 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20. Liu B, Yu X, Luo K, Yu Y, Yu XF. 2004. Influence of primate lentiviral Vif and proteasome inhibitors on human immunodeficiency virus type 1 virion packaging of APOBEC3G. J. Virol. 78:2072–2081 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21. Luo K, et al. 2005. Primate lentiviral virion infectivity factors are substrate receptors that assemble with cullin 5-E3 ligase through a HCCH motif to suppress APOBEC3G. Proc. Natl. Acad. Sci. U. S. A. 102:11444–11449 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22. Marcsisin SR, et al. 2011. On the solution conformation and dynamics of the HIV-1 viral infectivity factor. J. Mol. Biol. 410:1008–1022 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23. Mariani R, et al. 2003. Species-specific exclusion of APOBEC3G from HIV-1 virions by Vif. Cell 114:21–31 [DOI] [PubMed] [Google Scholar]
- 24. Mehle A, et al. 2004. Vif overcomes the innate antiviral activity of APOBEC3G by promoting its degradation in the ubiquitin-proteasome pathway. J. Biol. Chem. 279:7792–7798 [DOI] [PubMed] [Google Scholar]
- 25. Mehle A, Thomas ER, Rajendran KS, Gabuzda D. 2006. A zinc-binding region in Vif binds Cul5 and determines cullin selection. J. Biol. Chem. 281:17259–17265 [DOI] [PubMed] [Google Scholar]
- 26. Mehle A, et al. 2007. Identification of an APOBEC3G binding site in human immunodeficiency virus type 1 Vif and inhibitors of Vif-APOBEC3G binding. J. Virol. 81:13235–13241 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27. Mulder LC, Harari A, Simon V. 2008. Cytidine deamination induced HIV-1 drug resistance. Proc. Natl. Acad. Sci. U. S. A. 105:5501–5506 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28. Mulder LC, et al. 2010. Moderate influence of human APOBEC3F on HIV-1 replication in primary lymphocytes. J. Virol. 84:9613–9617 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29. OhAinle M, Kerns JA, Li MM, Malik HS, Emerman M. 2008. Antiretroelement activity of APOBEC3H was lost twice in recent human evolution. Cell Host Microbe 4:249–259 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30. Ooms M, Majdak S, Seibert CW, Harari A, Simon V. 2010. The localization of APOBEC3H variants in HIV-1 virions determines their antiviral activity. J. Virol. 84:7961–7969 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31. Paul I, Cui J, Maynard EL. 2006. Zinc binding to the HCCH motif of HIV-1 virion infectivity factor induces a conformational change that mediates protein-protein interactions. Proc. Natl. Acad. Sci. U. S. A. 103:18475–18480 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32. Pery E, Rajendran KS, Brazier AJ, Gabuzda D. 2009. Regulation of APOBEC3 proteins by a novel YXXL motif in human immunodeficiency virus type 1 Vif and simian immunodeficiency virus SIVagm Vif. J. Virol. 83:2374–2381 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33. Piantadosi A, Humes D, Chohan B, McClelland RS, Overbaugh J. 2009. Analysis of the percentage of human immunodeficiency virus type 1 sequences that are hypermutated and markers of disease progression in a longitudinal cohort, including one individual with a partially defective Vif. J. Virol. 83:7805–7814 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34. Russell RA, Pathak VK. 2009. Identification of Two Distinct HIV-1 Vif determinants critical for interactions with human APOBEC3G and APOBEC3F. J. Virol. 83:1992–2003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35. Schrofelbauer B, Senger T, Manning G, Landau NR. 2006. Mutational alteration of human immunodeficiency virus type 1 Vif allows for functional interaction with nonhuman primate APOBEC3G. J. Virol. 80:5984–5991 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36. Simon JH, Sheehy AM, Carpenter EA, Fouchier RA, Malim MH. 1999. Mutational analysis of the human immunodeficiency virus type 1 Vif protein. J. Virol. 73:2675–2681 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37. Simon V, et al. 2005. Natural variation in Vif: differential impact on APOBEC3G/3F and a potential role in HIV-1 diversification. PLoS Pathog. 1:e6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38. Smith JL, Bu W, Burdick RC, Pathak VK. 2009. Multiple ways of targeting APOBEC3-virion infectivity factor interactions for anti-HIV-1 drug development. Trends Pharmacol. Sci. 30:638–646 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39. Stanley BJ, et al. 2008. Structural insight into the human immunodeficiency virus Vif SOCS box and its role in human E3 ubiquitin ligase assembly. J. Virol. 82:8656–8663 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40. Stephens EB, Singh DK, Pacyniak E, McCormick C. 2001. Comparison of Vif sequences from diverse geographical isolates of HIV type 1 and SIV(cpz) identifies substitutions common to subtype C isolates and extensive variation in a proposed nuclear transport inhibition signal. AIDS Res. Hum. Retroviruses 17:169–177 [DOI] [PubMed] [Google Scholar]
- 41. Tan L, Sarkis PT, Wang T, Tian C, Yu XF. 2009. Sole copy of Z2-type human cytidine deaminase APOBEC3H has inhibitory activity against retrotransposons and HIV-1. FASEB J. 23:279–287 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42. Tian C, et al. 2006. Differential requirement for conserved tryptophans in human immunodeficiency virus type 1 Vif for the selective suppression of APOBEC3G and APOBEC3F. J. Virol. 80:3112–3115 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43. Wang X, et al. 2011. Analysis of human APOBEC3H haplotypes and anti-human immunodeficiency virus type 1 activity. J. Virol. 85:3142–3152 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44. Wichroski MJ, Ichiyama K, Rana TM. 2005. Analysis of HIV-1 viral infectivity factor-mediated proteasome-dependent depletion of APOBEC3G: correlating function and subcellular localization. J. Biol. Chem. 280:8387–8396 [DOI] [PubMed] [Google Scholar]
- 45. Wieland U, et al. 1994. In vivo genetic variability of the HIV-1 vif gene. Virology 203:43–51 [DOI] [PubMed] [Google Scholar]
- 46. Wieland U, et al. 1997. Diversity of the vif gene of human immunodeficiency virus type 1 in Uganda. J. Gen. Virol. 78:393–400 [DOI] [PubMed] [Google Scholar]
- 47. Xiao Z, Ehrlich E, Luo K, Xiong Y, Yu XF. 2007. Zinc chelation inhibits HIV Vif activity and liberates antiviral function of the cytidine deaminase APOBEC3G. FASEB J. 21:217–222 [DOI] [PubMed] [Google Scholar]
- 48. Xiao Z, et al. 2007. Characterization of a novel Cullin5 binding domain in HIV-1 Vif. J. Mol. Biol. 373:541–550 [DOI] [PubMed] [Google Scholar]
- 49. Yamashita T, Doi N, Adachi A, Nomaguchi M. 2008. Growth ability in simian cells of monkey cell-tropic HIV-1 is greatly affected by downstream region of the vif gene. J. Med. Invest. 55:236–240 [DOI] [PubMed] [Google Scholar]
- 50. Yu Y, Xiao Z, Ehrlich ES, Yu X, Yu XF. 2004. Selective assembly of HIV-1 Vif-Cul5-ElonginB-ElonginC E3 ubiquitin ligase complex through a novel SOCS box and upstream cysteines. Genes Dev. 18:2867–2872 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51. Zennou V, Perez-Caballero D, Gottlinger H, Bieniasz PD. 2004. APOBEC3G incorporation into human immunodeficiency virus type 1 particles. J. Virol. 78:12058–12061 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52. Zhen A, Wang T, Zhao K, Xiong Y, Yu XF. 2010. A single amino acid difference in human APOBEC3H variants determines HIV-1 Vif sensitivity. J. Virol. 84:1902–1911 [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.