ArticlePDF Available

Abstract and Figures

HIV-1 particle production is driven by the Gag precursor protein Pr55(Gag). Despite significant progress in defining both the viral and cellular determinants of HIV-1 assembly and release, the trafficking pathway used by Gag to reach its site of assembly in the infected cell remains to be elucidated. The Gag trafficking itinerary in primary monocyte-derived macrophages is especially poorly understood. To define the site of assembly and characterize the Gag trafficking pathway in this physiologically relevant cell type, we have made use of the biarsenical-tetracysteine system. A small tetracysteine tag was introduced near the C-terminus of the matrix domain of Gag. The insertion of the tag at this position did not interfere with Gag trafficking, virus assembly or release, particle infectivity, or the kinetics of virus replication. By using this in vivo detection system to visualize Gag trafficking in living macrophages, Gag was observed to accumulate both at the plasma membrane and in an apparently internal compartment that bears markers characteristic of late endosomes or multivesicular bodies. Significantly, the internal Gag rapidly translocated to the junction between the infected macrophages and uninfected T cells following macrophage/T-cell synapse formation. These data indicate that a population of Gag in infected macrophages remains sequestered internally and is presented to uninfected target cells at a virological synapse.
Content may be subject to copyright.
Real-Time Visualization of HIV-1 GAG Trafficking in
Infected Macrophages
Karine Gousset
1
, Sherimay D. Ablan
1
, Lori V. Coren
2
, Akira Ono
3
, Ferri Soheilian
4
, Kunio Nagashima
4
,
David E. Ott
2
, Eric O. Freed
1
*
1 Virus-Cell Interaction Section, HIV Drug Resistance Program, National Cancer Institute, Frederick, Maryland, United States of America, 2 AIDS Vaccine Program, SAIC-
Frederick, Inc., National Cancer Institute, Frederick, Maryland, United States of America, 3 Department of Microbiology and Immunology, University of Michigan Medical
School, Ann Arbor, Michigan, United States of America, 4 Image Analysis Laboratory, Advanced Technology Program, SAIC-Frederick, National Cancer Institute at
Frederick, Frederick, Maryland, United States of America
Abstract
HIV-1 particle production is driven by the Gag precursor protein Pr55
Gag
. Despite significant progress in defining both the
viral and cellular determinants of HIV-1 assembly and release, the trafficking pathway used by Gag to reach its site of
assembly in the infected cell remains to be elucidated. The Gag trafficking itinerary in primary monocyte-derived
macrophages is especially poorly understood. To define the site of assembly and characterize the Gag trafficking pathway in
this physiologically relevant cell type, we have made use of the biarsenical-tetracysteine system. A small tetracysteine tag
was introduced near the C-terminus of the matrix domain of Gag. The insertion of the tag at this position did not interfere
with Gag trafficking, virus assembly or release, particle infectivity, or the kinetics of virus replication. By using this in vivo
detection system to visualize Gag trafficking in living macrophages, Gag was observed to accumulate both at the plasma
membrane and in an apparently internal compartment that bears markers characteristic of late endosomes or multivesicular
bodies. Significantly, the internal Gag rapidly translocated to the junction between the infected macrophages and
uninfected T cells following macrophage/T-cell synapse formation. These data indicate that a population of Gag in infected
macrophages remains sequestered internally and is presented to uninfected target cells at a virological synapse.
Citation: Gousset K, Ablan SD, Coren LV, Ono A, Soheilian F, et al. (2008) Real-Time Visualization of HIV-1 GAG Trafficking in Infected Macrophages. PLoS
Pathog 4(3): e1000015. doi:10.1371/journal.ppat.1000015
Editor: Thomas J. Hope, Northwestern University, United States of America
Received October 2, 2007; Accepted Jan uary 30, 2008; Published March 7, 2008
This is an open-access article distributed under the terms of the Creative Commons Public Domain declaration which stipulates that, once placed in the public
domain, this work may be freely reproduced, distributed, transmitted, modified, built upon, or otherwise used by anyone for any lawful purpose.
Funding: This research was supported by the Intramural Research Program of the Center for Cancer Research, National Cancer Institute, NIH, by the Intramural
AIDS Targeted Antiviral Program, and with federal funds from the National Cancer Institute, NIH, under contract N01-CO-12400.
Competing Interests: The authors have declared that no competing interests exist.
* E-mail: efreed@nih.gov
Introduction
The human immunodeficiency virus type 1 (HIV-1) Gag
polyprotein precursor, Pr55
Gag
, plays an essential role in virus
assembly and release. Its expression alone is able to generate virus-
like particles (VLPs) [1,2]. All four domains of Pr55
Gag
–matrix
(MA), capsid (CA), nucleocapsid (NC) and p6–play important roles
in particle assembly and release [1,3]. The MA domain regulates
the association of Gag with the host cell plasma membrane (PM);
this membrane-binding activity is provided primarily by a myristic
acid moiety covalently attached to the N-terminus of MA and a
highly basic patch of amino acid residues that interacts with acidic
phospholipids, including phosphatidylinositol-(4,5)-bisphosphate
[PI4,5)P
2
] on the inner leaflet of the PM [4,5,6,7]. CA and NC
promote Gag-Gag interactions during assembly [8], in part
through the ability of NC to interact with nucleic acid [2,9].
Finally, the p6 domain of Gag stimulates virus release by
interacting with components of the cellular endosomal sorting
machinery [10,11,12].
Although significant progress has been made in elucidating the
viral and cellular factors necessary for Gag membrane binding,
Gag multimerization, and virus release, the subcellular location of
HIV-1 assembly has been the subject of controversy and the
itinerary of Gag trafficking to the site of assembly remains to be
defined. Mutational studies have shown that the viral determinants
for Gag targeting to the PM reside in the MA domain of Gag. A
large deletion in MA redirects HIV-1 assembly to the endoplasmic
reticulum [13,14], whereas point mutations, particularly in the
highly basic domain of MA, shift the site of assembly from the PM
to internal compartments [15,16,17] defined as late endosomes or
multivesicular bodies (MVBs) [18].
HIV-1 was long assumed to follow the classically defined ‘‘C-
type’’ pathway in which Gag assembly and release take place at
the PM [2]. This dogma was challenged by a number of studies
suggesting that HIV-1 assembly takes place in an endosomal
compartment and that particle release from the infected cell
follows the ‘‘exosomal’’ pathway in which virus-containing
endosomes fuse with the PM to release their contents
[19,20,21,22,23,24]. This endosomal model was then subsequently
contested by several studies showing PM-based HIV-1 assembly
and release [25,26,27,28,29]. The nature of the HIV-1 assembly
site in primary monocyte-derived macrophages (MDMs) has been
a matter of particular interest [30]. Early electron microscopy
(EM) observations in HIV-1-infected MDMs revealed an abun-
dance of virions assembling and budding into intracellular
vacuoles [31,32]. In later studies, it was observed that the virus-
containing internal compartments in MDMs bore markers
characteristic of late endosomes or MVBs; e.g., major histocom-
PLoS Pathogens | www.plospathogens.org 1 2008 | Volume 4 | Issue 3 | e1000015
patibility complex II (MHC II) and tetraspanins CD63, CD81, and
CD82 [18,33,34]. Furthermore, virions derived from MDMs
packaged late endosome/MVB markers, suggesting that these
virions originated from a late endosomal compartment [33,35,36].
In an intriguing refinement of the model that HIV-1 assembles in
MVBs in primary macrophages, it was demonstrated that at least
some of the virus-positive, ‘‘intracellular’’ structures in MDMs were
actually connected to the PM. These apparently internal structures
may therefore represent PM invaginations that are positive for
tetraspanin markers [37,38]. Elucidating the virus assembly
pathway in primary MDMs is highly significant since this cell type
represents one of the major targets for HIV-1 infection in vivo [39].
One of the difficulties in evaluating previous studies focused on
defining the Gag assembly/release pathway in MDMs is the
absence of live-cell imaging data in this cell type that allow the
trafficking of Gag to be visualized in real time. To this end, we
developed a system for visualizing in living cells the localization
and trafficking of Gag expressed in the context of a fully infectious
and replication-competent HIV-1 molecular clone. We used the
biarsenical-tetracysteine labeling method first described by Tsien
and colleagues [40,41,42]. This system is based on the insertion of
a small tetracysteine (TC) motif into a protein of interest. Cells
expressing the TC-tagged protein are treated with a membrane-
permeable biarsenical dye [e.g., green (FlAsH) or red (ReAsH)]
that fluoresces upon binding to the TC tag. The advantages of this
method are that the TC tag is very small and that labeling occurs
immediately upon binding of the dye to the TC tag. Recently, this
system was used to label Gag expressed from non-infectious clones
in HeLa, Mel Juso and Jurkat T cells [23,29]. We introduced the
TC tag near the C-terminus of the MA domain of Gag in the
context of the full-length infectious HIV-1 molecular clone pNL4-
3. Insertion of the TC tag had no significant effect on HIV-1 Gag
function. By using VSV-G-pseudotyped viruses, we were able to
infect and follow Gag trafficking in primary MDMs. Our data
indicate that in MDMs Gag accumulates both at the PM and in an
apparently internal MVB-like compartment. Although we ob-
tained no evidence for constitutive movement of the internal Gag
to the PM, or for internalization of the PM-localized Gag to
apparently internal structures, we observed rapid relocation of the
internal population of Gag to the site of cell-cell contact following
addition of susceptible T cells to the infected macrophage cultures.
These findings support a model whereby newly assembled virus
particles are sequestered in infected macrophages and then
efficiently presented to susceptible target cells following synapse
formation.
Results
Introduction of a TC tag near the C-terminus of the MA
domain of HIV-1 Gag does not disrupt virus replication,
assembly and release, or Ga g trafficking
The MA domain of HIV-1 Gag performs several important
functions in virus assembly and release [1]; however, deletion of a
number of C-terminal residues (amino acids 116-128) [43], or the
insertion of a Myc or green fluorescent protein (GFP) tag near the
C-terminus of MA [44] does not block virus assembly and release,
suggesting that the C-terminus of MA is relatively insensitive to
mutation. Thus, to facilitate the study of HIV-1 Gag trafficking,
we deleted codons 121–128 of MA and inserted a TC tag in the
full-length molecular clone pNL4-3 to generate pNL4-3/MA-TC
(Fig. 1A).
To determine the effects of the TC tag on virus replication, we
transfected the Jurkat T-cell line with WT pNL4-3 or with pNL4-
3/MA-TC and monitored virus replication over time by
measuring the levels of reverse transcriptase (RT) activity in the
medium (Fig. 1B). We observed that replication of NL4-3/MA-
TC was comparable to that of WT in Jurkat T-cells. To test the
replication of MA-TC in primary MDMs, the MA-TC tag was
introduced into the macrophage-tropic pNL4-3 derivative
pNL(AD8) [45,46]. Virus stocks were prepared and used to infect
MDMs. As indicated in Fig. 1B, the NL(AD8)/MA-TC virus
replicated with kinetics indistinguishable from those of WT
NL(AD8) in this physiologically relevant primary cell type.
The ability of MA-TC virus to replicate efficiently in both T-cell
lines and primary MDMs suggested that the insertion of the TC
tag near the C-terminus of MA does not affect HIV-1 assembly or
release. To test this directly, we transfected HeLa cells with WT
pNL4-3 or pNL4-3/MA-TC. One day posttransfection, the cells
were labeled for 5 minutes with or without FlAsH, washed with
ethanedithiol (EDT), and metabolically labeled for 2–3 hrs with
[
35
S]Met/Cys. Cell and virion lysates were prepared, immuno-
precipitated with anti-HIV immunoglobulin (HIV-Ig), subjected to
SDS-PAGE, and bands were quantitated by phosphorimager
analysis (Figure 1C) (see Materials and Methods). The results
indicated that insertion of the MA-TC tag had no significant effect
on virus particle production and that the FlAsH dye caused no
measurable disruption of HIV-1 particle production. We note that
insertion of the TC tag in MA resulted in increased labeling of the
MA protein with [
35
S]Met/Cys due to the additional Cys residues
(Fig. 1C). We also compared the single-cycle infectivity of WT and
MA-TC Gag in the TZM-bl indicator cell line [47] and observed
no effect of the MA-TC tag on virus infectivity (data not shown).
Together, these data demonstrate that the MA-TC tag does not
disrupt normal HIV-1 Gag function.
We previously reported a number of mutations within the MA
domain of Gag that alter normal HIV-1 Gag trafficking and
localization [15,16,18]. For example, mutation of the site of Gag
myristylation (1GA; [15]) results in a diffuse cytosolic Gag
localization. Mutations in the MA highly basic domain (e.g.,
29KE/31KE) retarget Gag to MVBs [16,18]. To validate further
the TC labeling approach, we sought to confirm that the effect of
these mutations on Gag localization in the context of otherwise
WT Gag would be recapitulated in the context of MA-TC Gag.
We introduced the 1GA and 29KE/31KE MA mutations into
Author Summary
The viral Gag protein is both necessary and sufficient for
the assembly of a new generation of virus particles. There
has been a significant amount of debate in recent years
regarding the site in the cell at which HIV-1 assembly takes
place. Of particular interest has been the site of assembly
in macrophages, a cell type that serves as an important
target for HIV-1 infection in vivo. In this study, we examine
the site of Gag localization and virus assembly in primary
human macrophages in living cells by using biarsenical
dyes that become fluorescent when they bind a small
target sequence introduced into HIV-1 Gag. We observe
Gag localization both at the plasma membrane and in an
apparently internal compartment that bears markers
characteristic of multivesicular bodies (MVBs). Significantly,
when infected macrophages are cocultured with uninfect-
ed T cells, the apparently internal Gag moves rapidly to the
contact site, or synapse, between the macrophage and the
T cell. These findings support the hypothesis that infected
macrophages sequester assembled HIV-1 particles in an
internal compartment and that these particles move to
synapses where cell–cell transmission can occur.
HIV-1 Gag in Macrophages
PLoS Pathogens | www.plospathogens.org 2 2008 | Volume 4 | Issue 3 | e1000015
MA-TC and analyzed Gag localization within cells using a rapid
FlAsH labeling method. Transfected HeLa cells were labeled for
5 min with FlAsH and washed for 20 min in EDT. Cells were then
fixed and either mounted or processed further for antibody
labeling. Similar to our previous results obtained with antibody
labeling [16,48], MA-TC Gag was found primarily in a punctate
pattern at the cell surface. In contrast, MA-TC/1GA was diffusely
localized throughout the cytosol and MA-TC/29KE/31KE was
found in internal compartments (Fig. 2, top). We previously
observed that the internal compartment to which 29KE/31KE
localizes in HeLa cells is positive for the MVB marker CD63 [18].
To verify that this was also the case in the context of MA-TC Gag,
we examined the colocalization of the 29KE/31KE-TC mutant,
labeled with ReAsH, with CD63 in transfected HeLa cells. We
observed nearly complete colocalization between 29KE/31KE
Gag and CD63 (Fig. S1). These results confirm the biochemical
experiments indicating that the addition of the TC tag had no
effect on Gag trafficking, assembly, or release in HeLa cells.
Since our goal was to visualize Gag trafficking in physiologically
relevant primary cells, we analyzed the localization pattern of MA-
TC Gag in infected MDMs. Cells were infected with VSV-G-
pseudotyped virus stocks obtained from transfected 293T cells.
Infected MDMs were then labeled with FlAsH and fixed 24 to
72 hours post-infection. Infection efficiencies, as determined by
Gag staining, typically ranged between 2 and 10%. MA-TC Gag
localized both to the PM and to an apparently internal
compartment (Fig. 2, bottom). In agreement with our previous
results obtained by antibody labeling [16,18,49], the localization of
MA-TC-derived 1GA and 29KE/31KE Gag in MDMs was
similar to that observed in HeLa cells: 1GA-TC was diffusely
distributed throughout the cytoplasm, and 29KE/31KE-TC was
almost exclusively found in apparently internal compartments
(Fig. 2, bottom). To provide a clearer visualization of the internal
localization of 29KE/31KE-TC Gag in MDM, we obtained a z-
series reconstruction by using the Maximum Intensity Projection
mode from the image processing software OsiriX (Video S1). The
results presented in Fig. 2 demonstrate that introduction of the TC
tag near the C-terminus of MA (MA-TC) allows HIV-1 Gag to be
readily visualized in infected primary MDMs at early time points
postinfection.
In MDM, Gag localizes to both the PM and to an internal,
tetraspanin-positive compar tment
We and others have previously reported that the apparently
internal vesicles to which HIV-1 Gag localizes in MDMs bear
Figure 1. The MA-TC tag does not affect Gag function. (A) Schematic diagram of HIV-1 Gag indicating the position of TC tag insertion. The
amino acid sequence of the TC tag is shaded. (B) Replication kinetics of WT HIV-1 vs. the MA-TC derivative in the Jurkat T-cell line and primary MDM.
Jurkat and MDM experiments were performed with pNL4-3 and pNL(AD8) molecular clones, respectively. Media were obtained every two days for RT
analysis. (C) Virus release efficiency of WT vs. MA-TC. HeLa cells were transfected with WT pNL4-3 or pNL4-3/MA-TC plasmids. Transfected cells were
labeled with FlAsH or DMSO (control) for 5 min 24–48 hrs posttransfection and were washed for 20 min with 300 mM EDT/PBS. The cells were then
metabolically labeled with [
35
S]Met/Cys for 2 hrs. Released virions were pelleted by ultracentrifugation, and both cell and virus lysates were
immunoprecipitated with HIV-Ig and subjected to SDS-PAGE. Bands were quantified using a phosphorimager. + /2 SD, n = 3.
doi:10.1371/journal.ppat.1000015.g001
HIV-1 Gag in Macrophages
PLoS Pathogens | www.plospathogens.org 3 2008 | Volume 4 | Issue 3 | e1000015
tetraspanin markers, suggesting that they are MVBs or MVB-like
structures [18,33,34]. To define the site of Gag localization in
MDMs at early time points postinfection using TC-tagged Gag,
we infected MDMs and examined the localization of Gag and
tetraspanins (CD63 and CD81) at 20 hrs postinfection. As
previously observed with fully WT Gag [18], MA-TC Gag
displayed a localization pattern that partially overlapped with that
of CD63 (Fig. 3A). The colocalization pattern in these cells was
very heterogeneous, with some cells displaying a high degree of
Gag/CD63 colocalization (Fig. 3A, top panels) and other cells
showing a lower level of colocalization (Fig. 3A, lower panels).
29KE/31KE-TC Gag also overlapped with a subset of CD63 in
infected MDMs (Fig. 3B). Both MA-TC (Fig. 3C) and 29KE/
31KE-TC Gag (data not shown) showed much more extensive
colocalization with CD81 than with CD63. We note that some
cells displayed a high level of Gag and CD81 costaining at the PM
(Fig. 3C, lower panel), consistent with HIV-1 assembly occurring
in tetraspanin-enriched microdomains at the cell surface [19,50].
To quantitatively compare the degree of Gag/CD63 vs. Gag/
CD81 colocalization in MDMs, we measured the Pearson
correlation coefficient (R) values (see Materials and Methods) for
these two sets of colocalizing proteins in a total of 75 cells. The
results confirmed the higher degree of Gag/CD81 compared to
Gag/CD63 colocalization (Fig. S2). We observed that 71% of cells
displayed a Gag/CD63 R-value of ,0.6, whereas 91% of cells
showed a Gag/CD81 R-value of .0.6 (Fig. S2).
As indicated in Fig. 3, WT Gag was localized both to an internal
tetraspanin-positive compartment and to the PM in infected
MDMs. Very few cells showed exclusively PM staining; instead,
the vast majority of cells showed either an internal localization or
both PM and internal staining. To determine whether the
distribution changed over time, we classified cells as displaying
uniquely PM, intracellular, or both PM and intracellular Gag
localization at 20, 24, 48, 72, and 96 hrs postinfection. The
percentage of cells within these three categories remained
essentially unchanged over time (Fig. 4A).
To visualize Gag movement in living MDMs, cells were infected
with MA-TC virions pseudotyped with VSV-G and were labeled
for 5 min with FlAsH 24 to 72 hrs post-infection. After washing,
labeled MDMs were immediately placed in a microscope chamber
(37uC/5% CO
2
) and imaged over time. Interestingly, no clear
movement of Gag between PM and apparently intracellular
compartments was observed during the time course (Fig. 4B); i.e.,
no obvious internalization of Gag from the PM was visualized, nor
was there clear movement of internal Gag puncta to the PM.
These results suggest that Gag can assemble both at the PM and in
internal compartments in infected MDMs. As stated in the
Materials and Methods, prior to 20 hrs post-infection we were not
able to definitively distinguish between specific Gag staining and
the diffuse, low-level background.
Gag and CD81 accumulate at sites of cell-to-cell contact
During the course of our analyses, we frequently observed
concentrated Gag staining at the contact sites formed between
infected and uninfected MDMs (Fig. 5A). 3D z-stack reconstruc-
tions illustrating this phenomenon are presented in Figs. S3A and
Video S2). These Gag-enriched cell-cell junctions also displayed a
high degree of staining for the tetraspanin markers CD81 and
CD82 (Fig. 5B and data not shown). Analogous junctions have
been reported to form between HIV-1-treated dendritic cells and
T-cells; because these junctions bear markers (e.g., tetraspanins
and adhesion molecules) found at immunological synapses
[51,52,53] they have been named ‘‘infectious’’ or ‘‘virological’’
synapses [54,55,56,57,58]. A concentration of budding and
released virions was also observed in the vicinity of cell-cell
contact sites by transmission electron microscopy (EM) (Fig. 5C).
To quantify the localization of Gag at the synapse observed in our
EM analysis, we counted the number of virus particles and
budding structures at synapse vs. non-synapse regions of the
plasma membrane. More than 60 cells were scored for this
analysis. The results indicated a markedly (5-6-fold) higher density
Figure 2. The MA-TC tag does not affect Gag localization. HeLa cells (top panel) were transfected with pNL4-3/MA-TC or 1GA or 29KE/31KE
derivatives. Cells were labeled with FlAsH for 5 min at 37uC, washed, fixed in 3.7% formaldehyde, and examined microscopically. MDMs (bottom
panel) were infected with VSV-G- pseudotyped virus stocks that transduced NL4-3/MA-TC or 1GA or 29KE/31KE derivatives (see Materials and
Methods). Infected cells were labeled and fixed as described above. Scale bars = 30 mm. For a 3D z-series of the 29KE/31KE-TC mutant in MDM, see
Video S1.
doi:10.1371/journal.ppat.1000015.g002
HIV-1 Gag in Macrophages
PLoS Pathogens | www.plospathogens.org 4 2008 | Volume 4 | Issue 3 | e1000015
of particles and budding events at synapse vs. non-synapse regions
of the cell surface, consistent with the immunofluorescence data
presented above.
To extend the analysis of Gag concentration at the cell-cell
synapse to include junctions formed between infected MDMs and
uninfected T-cells, we performed the following analysis: infected
MDMs were labeled with FlAsH and then incubated at 37uC for
2 hours with Jurkat T-cells. The cells were then fixed and, when
necessary, labeled with anti-CD81 antibodies. Gag was frequently
detected at the synapses between infected MDMs and uninfected
Jurkat T-cells (Fig. 6A). 3D z-stack reconstructions are provided in
Videos S3 and S4. Furthermore, as we observed for MDM/MDM
junctions, MDM/T-cell synapses also displayed a high degree of
colocalization between Gag and tetraspanin markers (Fig. 6B). We
also observed that Gag concentrated at synapses formed between
infected MDMs and primary T cells (data not shown). Overall,
these data show that HIV-1 Gag, along with CD81, are recruited
to the synapses formed between infected macrophages and
uninfected macrophages or T-cells. To quantify the concentration
of Gag at the synapse, we used the ImageJ software to determine
the pixel intensity for Gag staining at the MDM/MDM and
MDM/T-cell synapses compared to the overall pixel intensity in
each infected cell. The results confirmed a high degree of Gag
concentration at cell-cell junctions, with approximately 80% of the
total Gag signal localized to the synapse (Figs. 6C, S3A).
Movement of Gag to the synapse in MDM is not Env-
dependent but is disrupted by mutations in MA
To analyze further the process of Gag recruitment to the
synapse in infected MDMs, we determined whether Gag was
recruited to cell-cell junctions in the context of proviral clones
carrying additional mutations. We first examined the localization
of Gag in the absence of Env expression by using the Env(-) MA-
TC mutant KFS/MA-TC. Examining a possible role for Env in
Gag recruitment to the MDM synapse was of interest as it has
been reported that Env is required for synapse formation between
infected and uninfected T cells [59] and also plays a role in the
formation of filopodial bridges that can facilitate transfer of
retroviruses between cells [60]. In contrast to these prior findings
in non-monocytic cell types, we observed that Gag was efficiently
localized to both MDM/MDM and MDM/T-cell synapses in the
absence of Env expression (Fig. 7A and B). This concentration of
Gag to the synapse was quantified as described above, confirming
the high degree of localization of Gag to the cell-cell junction
independent of Env expression (Fig. 7C, S3B). 3D z-stack
reconstructions are provided in Video S5. The data indicated no
statistically significant difference between Gag concentration at the
MDM/T-cell vs, MDM/MDM synapse, or in the presence or
absence of Env expression (compare Figs. 6C and 7C).
We previously reported that mutations in the highly basic
domain of MA (e.g., 29KE/31KE) redirect Gag to MVBs [18].
Here, we observed that in MDMs the 29KE/31KE-TC mutant
displayed nearly complete localization to an apparently internal
compartment that stained positive for CD63 and CD81 (Figs. 2
and 3 and data not shown). In contrast, MA-TC Gag displayed a
mix of PM and internal staining (e.g., Fig. 2–4). It was therefore of
interest to examine whether 29KE/31KE-TC Gag could
redistribute from its normally internal site of localization to the
cell surface upon synapse formation. Interestingly, we observed
that in contrast to MA-TC Gag, 29KE/31KE-TC Gag did not
relocalize to either MDM/MDM (Fig. 8; Video S6) or MDM-T-
cell (data not shown) synapse. Instead, in cells expressing 29KE/
31KE-TC Gag, both Gag and CD81 remained deep within the
infected cell (Fig. 8). In four independent experiments with 29KE/
31KE-TC, Gag accumulation was never observed at the synapse.
These data suggest the possibility that the apparently internal
compartments to which WT and 29KE/31KE Gag localize are
distinct.
Real-time visualization of Gag movement to the
infectious synapse in living MDM
The data presented above using fixed infected cells and EM
techniques demonstrate the accumulation of Gag and virus
particles at the junction between infected MDMs and uninfected
MDMs or T-cells. To visualize the movement of Gag to the cell-
cell contact site, we used FlAsH labeling and live-cell imaging in
infected MDMs. For these experiments, infected MDMs were
labeled with FlAsH for 5 minutes 24 to 72 hours post-infection,
washed, and imaged over time. When visualizing MDM/T-cell
junctions, Jurkat T-cells were added to the infected cells post-
Figure 3. Gag in MDMs colocalizes with tetraspanins at the cell
surface and in apparently interna l compartment s. MDMs
infected with VSV-G-pseudotyped NL4-3/MA-TC (A and C) or NL4-3/
29KE/31KE-TC (B) were labeled with FlAsH (green) 20 hours post-
infection, fixed, and further labeled with antibody against CD63 (red) (A
and B) or CD81 (red) (C). The merged images of Gag-TC and CD63/CD81
are shown on the right, with yellow indicating colocalization between
Gag and CD63/CD81. R = Pearson coefficient of correlation. Scale bars:
panel A, 40 mm; panel B, 30 mm; panel C; 30 mm top, 20 mm bottom.
doi:10.1371/journal.ppat.1000015.g003
HIV-1 Gag in Macrophages
PLoS Pathogens | www.plospathogens.org 5 2008 | Volume 4 | Issue 3 | e1000015
FlAsH labeling and imaged under the same conditions. After
addition of the Jurkat cells, incubation periods of approximately
30–60 min were required for stable MDM/T-cell synapses to
form. As mentioned above, we observed no clear evidence of
movement of apparently internal Gag to the PM, or vice versa, in
MDMs not actively engaged in cell-cell contact. Interestingly,
however, upon addition of Jurkat T-cells to the infected MDM
cultures, we observed rapid movement of apparently internal Gag
to the MDM/T-cell synapse. The infected macrophage (‘‘M1’’) in
Fig. 9A is surrounded by uninfected macrophages (e.g., ‘‘M2’’) and
Jurkat T-cells (‘‘T1’’ and ‘‘T2’’). In this particular time course,
40 min after adding Jurkat T cells to the MDMs (t = 0 min), Gag
has already accumulated at the contact site between M1 and T1.
Gag-containing compartments are also rapidly recruited to the site
of M1/T2 contact. Movement of other Gag-containing compart-
ments toward the site of MDM-MDM (M1/M2) contact can be
observed starting at time t = 25 min and is complete at t = 40 min.
These data demonstrate that Gag present in internal compart-
ments can be rapidly redistributed to the site of contact with
uninfected cells. After its movement to the MDM/MDM synapse,
Gag can be seen moving along the surface of macrophage M2
(e.g., at 45 and 50 min). A movie of Gag movement to the synapse
can be viewed at Video S7. Intriguingly, we frequently observed
an apparent preference for MDM/T-cell synapse formation at
sites close to high levels of Gag concentration (Fig. 9B). In this
gallery, time t = 0 represents cells 90 min post-FlAsH labeling and
70 min after addition of T-cells. One of the surrounding T-cells
(‘‘T ‘‘) at time t = 0 min extends on top of the infected MDM
toward the site of Gag accumulation. This resulted in the
movement and attachment of the T-cell with the infected MDM
near the site of Gag accumulation.
Discussion
Most studies that have examined HIV-1 Gag trafficking have
used non-infectious constructs in which codon-optimized Gag is
fused to fluorescent proteins such as green or red fluorescent
protein (GFP or RFP). Although these studies provided important
insights, disadvantages of using GFP and its derivatives in protein
trafficking analyses include the large size of the fluorescent protein
and the fact that achieving their fluorescent state requires time-
dependent chromophore maturation [61,62]. We have also
observed that Gag expressed from some codon-optimized
constructs assembles relatively inefficiently and forms perinuclear
cytosolic aggregates not typically observed with WT Gag
(unpublished results). In this study, we describe the application
of the biarsenical labeling system to visualize HIV-1 Gag
trafficking in primary MDMs. We show here that the TC tag
Figure 4. Gag localization in MDMs remains stable over time. (A) MDMs were labeled with FlAsH 20, 24, 48, 72, and 96 hrs after infection with
VSV-G-pseudotyped NL4-3/MA-TC. At each time point, between 25 and 65 cells were categorized as having PM, intracellular (IC), or both
PM+intracellular (PM+IC) Gag localization. (B) Live-cell analysis. MDMs infected with VSV-G-pseudotyped NL4-3/MA-TC virus were labeled with FlAsH
48 hrs postinfection, and immediately examined microscopically on the stage in a closed chamber (37uC/5% CO
2
). Time (t) represents time in min
after FlAsH labeling. Scale bar = 30 mm.
doi:10.1371/journal.ppat.1000015.g004
HIV-1 Gag in Macrophages
PLoS Pathogens | www.plospathogens.org 6 2008 | Volume 4 | Issue 3 | e1000015
Figure 5. Gag accumulates at the MDM/MDM synapse. Cells were infected with VSV-G-pseudotyped NL4-3/MA-TC virus, labeled with FlAsH
24–48 hrs post-infection, and fixed in 3.7% formaldehyde. (A) Gag accumulates at the synapse between MDMs. Individual macrophages are labeled
‘‘M’’. Scale bar = 15 mm. (B) After FlAsH labeling, cells were fixed and stained with anti-CD81 antibody. The extensive overlap between Gag and CD81
at the MDM/MDM synapse is visualized as yellow in the merged panels. Scale bar = 30 mm. Far-right panels in (A and B) provide quantification of the
Gag signals. The Surface Plot analyzing tool from the ImageJ software was used to obtain a three-dimensional graph of the pixel intensities in
grayscale. The x- and y- axes represent the length of the region analyzed in pixels, and the z-axis represents the pixel intensity of the Gag signal. (C)
EM analysis of VLPs at the synapse. Infected MDMs were fixed with 2% glutaraldehyde and processed for EM. Fully assembled VLPs can be visualized
at the MDM/MDM junctions. Red boxes represent regions that are enlarged in the adjacent panels.
doi:10.1371/journal.ppat.1000015.g005
HIV-1 Gag in Macrophages
PLoS Pathogens | www.plospathogens.org 7 2008 | Volume 4 | Issue 3 | e1000015
that serves as the binding site for the biarsenical dye FlAsH is
remarkably well tolerated with respect to preserving Gag function
when introduced near the C-terminus of the MA domain. The
MA-TC tagged Gag produces virus particles with WT efficiency,
and these particles are fully infectious in both single-cycle assays
and in spreading infections. The MA-TC Gag can be readily
delivered to primary cells as a VSV-G pseudotype. Overall, this
system provides a rapid and efficient method for observing WT
Gag trafficking in living cells.
The biarsenical labeling system employed in this study allowed
us to examine Gag localization after ,20 hrs postinfection, the
earliest time point at which Gag expression could be readily and
consistently visualized. At this early time point, we observed a mix
of PM and apparently internal Gag staining. Between 20 and
Figure 6. Gag accumulates at the MDM/T-cell synapse. (A) Infected MDMs were labeled with FlAsH and Jurkat cells were added to the cultures.
One to two hours after addition of the T-cells, the co-cultures were fixed with 3.7% formaldehyde and imaged or (B) stained with anti-CD81 antibody.
In panel B, Gag/CD81 colocalization is indicated as yellow in merge. M = MDMs; T = Jurkat T-cells. Scale bars in panels A and B = 15 mm. (C)
Quantification of Gag concentration at the synapse. The microscopy images were opened in ImageJ and a plot profile was obtained. A column
average plot was generated, in which the x-axis represents the horizontal distance through the image and the z-axis the vertically averaged pixel
intensity. The % Gag at the synapse = (pixel intensity of Gag signal at synapse)/(total intensity of Gag in the cell+synapse)6100. N values indicate the
number of cells analyzed in each data set. For more information on how the Gag quantification was performed, see Fig. S3A.
doi:10.1371/journal.ppat.1000015.g006
HIV-1 Gag in Macrophages
PLoS Pathogens | www.plospathogens.org 8 2008 | Volume 4 | Issue 3 | e1000015
96 hrs postinfection, we did not observe a shift in the percentage of
cells displaying PM, intracellular, or PM+intracellular staining
(Fig. 4) nor did we observe a time-dependent accumulation of
internal Gag. These results differ from those of a recent study in
which intracellular Gag-GFP staining increased over time [27].
We note that the TC-tagged Gag in the current study is fully
functional for particle assembly and release and produces
infectious virions. Furthermore, our TC-tagged Gag is expressed
in the context of a full-length molecular clone that encodes all the
HIV-1 accessory proteins including Vpu. Indeed, elimination of
Vpu expression led to a time-dependent accumulation of Gag in
internal compartments (unpublished results), consistent with recent
reports [26,28]. In agreement with the study of Jouvenet et al.
[27], we did not observe an effect on virus release of treating
infected cells with U18666A, a drug that arrests endosome motility
(unpublished results). This observation supports the hypothesis
that release of HIV-1 in MDMs occurs from PM-assembled VLPs.
Although the FlAsH method can be accompanied by high
background staining, we observed that this problem is largely
mitigated by using very brief labeling periods. We also observed
Figure 7. Gag recruitment to the synapse in MDMs is Env-independent. (A) MDMs were infected with VSV-G-pseudotyped NL4-3/KFS/MA-TC
virus (which is defective for HIV-1 Env expression) and were labeled with FlAsH 24–48 hrs postinfection. Boxes indicate regions enlarged on the right.
Scale bars = 15 mm. (B) MDMs were infected as in (A), and Jurkat T cells were added after FlAsH labeling. Far-right panels in (A) and (B) provide
quantification of the Gag signals, determined as indicated in the Fig. 5 legend. Scale bars = 15 mm. (C) Quantification of Gag concentration at the
synapse, determined as described in the Fig. 6C legend. For more information on how the Gag quantification was performed, see Fig. S3B.
doi:10.1371/journal.ppat.1000015.g007
HIV-1 Gag in Macrophages
PLoS Pathogens | www.plospathogens.org 9 2008 | Volume 4 | Issue 3 | e1000015
that background staining in MDM is less evident in MDM than in
HeLa cells.
The most significant finding in this study is the visualization of
apparently internal Gag moving to the site of cell-cell contact after
synapse formation with uninfected T cells. It therefore appears
that the tetraspanin-rich, apparently internal compartment in
which HIV-1 assembles in MDMs can serve as a storage
compartment for rapid presentation of virus particles at cell-cell
junctions. These findings have clear implications for HIV-1
transmission from MDMs. In this regard, it is interesting to note
that the virions in these internal vesicles reportedly remain
infectious for weeks postinfection [63] and that virus transmission
between infected MDMs and T cells is extremely rapid [64]. In
several respects, our observations with infected MDMs are similar
to those made with dendritic cells treated with HIV-1. Binding of
HIV-1 virions to dendritic cells can lead to transfer of virus to
uninfected T-cells through the formation of a virological or
infectious dendritic cell/T-cell synapse without the dendritic cell
itself being productively infected [55,56,57,58,65,66]. Virions
bound to the dendritic cell are reportedly internalized into an
internal compartment that is strongly positive for CD81 but only
weakly positive for CD63. Synapse formation induces the
redistribution of virus particles and CD81 to the site of cell-cell
contact, presumably facilitating transfer of virus to the T cell. The
internal virus-containing compartment in dendritic cells is weakly
acidic, as also reported for the virus-positive compartment in
MDMs [67]. Thus, it appears that HIV-1 has evolved to subvert a
pathway in both MDMs and dendritic cells that allows infectious
virus particles to be retained in an apparently internal compart-
ment and then redistributed to the cell surface following infectious
synapse formation. Transfer of HIV-1 between T cells also
involves the formation of a synapse that bears tetraspanin markers
[68]; however, there is currently no evidence for long-term
retention of infectious virus particles within an internal compart-
ment in T cells. Interestingly, whereas the generation of T-cell/T-
cell infectious synapses [59] and formation of cell-cell filopodial
bridges that allow intercellular transfer of virus particles [60] have
been reported to require Env expression, we found no such Env
dependence for Gag translocation to the synapse formed between
MDMs and T cells (Fig. 7A).
We previously reported that mutations in the highly basic
domain of MA (e.g. 29KE/31KE) induce a shift in Gag
localization in HeLa cells and T cells from PM to MVBs [18].
In MDMs, both WT and 29KE/31KE Gag localize to an
apparently internal, tetraspanin-positive compartment [18]. This
finding is confirmed here (e.g., Fig. 2 and 8). Interestingly, while
WT MA-TC Gag rapidly translocates to the MDM/T-cell
junction after synapse formation, we did not observe significant
movement of 29KE/31KE-TC Gag to the MDM/T-cell synapse
(Fig. 8). These results imply that WT and 29KE/31KE Gag
localize to distinct tetraspanin-positive compartments in MDMs. A
possible interpretation of these observations is that 29KE/31KE
Gag localizes to ‘‘true’’ MVBs which do not move to the synapse,
whereas WT Gag localizes to a compartment that is apparently
internal but is connected to the PM [37,38]. It is this PM-
connected tetraspanin-positive compartment that undergoes a shift
in localization following synapse formation, thereby allowing virus
particle movement to the site of cell-cell contact. Surprisingly, we
frequently observed that T-cells made contact with regions of the
MDM PM under which Gag was concentrated, and in many cases
the T-cells formed pseudopodia to contact this site (e.g., Fig. 9B).
These results imply that the T cell can ‘‘sense’’ regions of the PM
that overlie the putative invaginations in which assembled virus
particles are concentrated. These regions of the PM may be
enriched in lipid rafts and/or tetraspanin-enriched microdomains.
These observations are somewhat reminiscent of previous studies
on the recruitment of uninfected T cells into infected cell syncytia
[69]. A future challenge will be to characterize in greater detail the
membrane composition at the site of MDM/T-cell contact and
elucidate the signals that induce the movement of the newly
assembled, internally sequestered virus particles to the infectious
synapse.
Figure 8. Gag recruitment to the synapse in MDMs is blocked by the 29KE/31KE MA mutations. MDMs were infected with VSV-G-
pseudotyped NL4-3/29KE/31KE-TC virus, labeled with FlAsH 24–48 hrs postinfection, fixed, and stained with anti-CD81 antibody. Gag/CD81
colocalization is indicated as yellow in merge. Scale bars = 30 mm. Far-right panels show distribution of Gag signal; plots were obtained as described
in the Fig. 5 legend. Note the centrally located (non-synapse) concentration of Gag in these cells expressing the 29KE/31KE MA mutant.
doi:10.1371/journal.ppat.1000015.g008
HIV-1 Gag in Macrophages
PLoS Pathogens | www.plospathogens.org 10 2008 | Volume 4 | Issue 3 | e1000015
Materials and Methods
Plasmids and preparation of virus stocks
Plasmids pNL4-3/MA-TC and pNL(AD8)/MA-TC were
constructed as follows: for pNL4-3/MA-TC, nucleotides 1250–
1273 (encoding DTGNNSQV Gag codons 121–128) were deleted
in the MA-coding region of the full-length HIV-1 molecular clone
pNL4-3 [70] and the TC tag GSMPCCPGCCGSM was inserted
in its place using overlap-extension PCR [71]. The MDM-tropic
pNL(AD8)/MA-TC clone was constructed by exchanging the
EcoRI-XhoI fragment of pNL4-3/MA-TC with that from the
CCR5-tropic clone pNL(AD8) [45]. Construction of molecular
clones expressing pNL4-3 MA mutants 1GA and 29KE/31KE
was described previously [15,16]. The molecular clones pNL4-3/
1GA-TC and pNL4-3/29KE/31/KE-TC were constructed by
exchanging the BssHII-SphI fragments of pNL4-3/1GA or pNL4-
Figure 9. Real-time trafficking of Gag to the synapse in infected MDMs. (A) MDMs infected with VSV-G-pseudotyped NL4-3/MA-TC virus
were labeled with FlAsH 48 hrs postinfection. Jurkat T-cells were added to the FlAsH-labeled cultures and incubated on the microscope stage in a
closed chamber (37uC /5% CO
2
). Time t = 0 min is 40 min (gallery A) or 70 min (gallery B) after addition of Jurkat cells. M1 = infected MDM; M2 = non-
infected MDM; T1 and T2 are Jurkat T-cells. Scale bars = 20 mm in A and 40 mm in B. For a movie of Gag movement to the synapse, derived from the
experiment presented in Fig. 9A, see Video S7.
doi:10.1371/journal.ppat.1000015.g009
HIV-1 Gag in Macrophages
PLoS Pathogens | www.plospathogens.org 11 2008 | Volume 4 | Issue 3 | e1000015
3/29KE/31KE with the corresponding fragments from MA-TC.
To construct the Env(-) pNL4-3 construct, pNL4-3/KFS-TC, we
exchanged the EcoRI-XhoI fragment from the Env(-) molecular
clone pNL4-3/KFS [72] with the corresponding fragment from
pNL4-3/MA-TC. Finally, we constructed pNL4-3/Vpu(-)/MA-
TC by replacing the BssHII-EcoRI fragment from Vpu-DEL-1
[73] (kindly provided by K. Strebel), with the corresponding
fragment from pNL4-3/MA-TC. VSV-G-pseudotyped virus
stocks were prepared by transfecting 293T cells with the Gag/
Pol expression vector pCMVNLGagPolRRE [74], the VSV-G
expression vector pHCMV-G [75], and the indicated HIV-1
molecular clones by using Lipofectamine 2000 (Invitrogen),
according to the manufacturer’s protocol.
Cells, transfections, and infections
HeLa and Jurkat T cells were cultured as previously described
[15]. MDMs were prepared by culturing elutriated monocytes
[45] in RPMI-1640 medium, supplemented with 10% fetal bovine
serum, for 5 to 7 days on ultra-low attachment plates (Costar).
HeLa cells were transfected by using the calcium phosphate
method, as previously described [15]. Jurkat T-cells were
transfected by using the DEAE-dextran procedure as previously
reported [15]. Infection of MDMs was performed as follows:
MDMs were detached from the ultra-low attachment plates
(Fisher Scientific, Pittsburgh, PA) and plated onto tissue culture
dishes or microscope culture chambers Fisher Scientific, Pitts-
burgh, PA). Virus stocks, pseudotyped with VSV-G, were
incubated with MDMs for 5-6 hours. 2610
6
counts/minute
(cpm) of reverse transcriptase (RT) activity was used per well of
4-well Nunc chambers, 10
6
RT cpm/well for 8-well Nunc-
chambers, and 4610
6
RT cpm/well for 6-well plates.
Virus replication and infectivity assays
Virus replication assays in the Jurkat T-cell line were performed
as previously described [15]. Briefly, Jurkat cells were transfected
in parallel with WT pNL4-3 or MA-TC using the DEAE-dextran
method. Cells were split 1:3 every two days and an aliquot of
medium was reserved at each time point for RT assay [76].
MDMs in 6-well plates were infected with 2610
6
RT cpm/well
with WT pNL(AD8) or MA-TC(AD8) virus stocks. Medium in the
infected MDM cultures was changed every two days and an
aliquot was reserved for RT activity. For single-cycle infectivity
assays, 4610
5
HeLa-derived TZM-bl cells [47] (obtained from J.
Kappes through the NIH AIDS Research and References Reagent
Program) per well were infected with 2610
5
RT cpm virus stocks.
Infection efficiency was determined by measuring luciferase
activity 2 days post-infection, as described previously [77].
Biarsenical labeling
Adherent cells cultured in Lab-Tek chamber slides (Nunc) or 6-
well plates were labeled 24–72 hours posttransfection/infection.
All labeling steps were performed at 37uC in the dark. The cells
were washed twice with Opti-MEM I (Invitrogen, Carlsbad). For
each experiment, biarsenical labeling solutions were freshly
prepared immediately prior to use. Wash solutions of 300 mM
and 100 mM 1,2-ethaneditiol (EDT) (Aldrich Chemical Company,
Inc., Milwaukee) were prepared in phosphate-buffered saline
(PBS) and 0.2 mM Lumio Green (FlAsH) was prepared in
dimethyl sulfoxide (DMSO) (Sigma-Aldrich, Inc., St Louis). Before
labeling, 2 ml of 1 mM EDT was mixed quickly with 4.7 mlof
0.2 mM FlAsH or Lumio Red (ReAsH) and immediately added to
400 ml Opti-MEM I. This solution was added to cells, which were
incubated for 5 min at 37uC. After the 5 min biarsenical labeling,
cells were washed with 300 mM EDT/PBS for 8 min and 100 mM
EDT/PBS for 10 min at 37uC. Cells were then washed further 3X
with PBS and either fixed with 3.7% formaldehyde prior to
antibody labeling or incubated with Opti-MEM I for live cell
imaging or addition of Jurkat T cells. The levels of background in
MDMs and in Hela cells were greatly reduced with the addition of
EDT in our labeling solutions. We also observed that keeping the
biarsenical labeling time short (5 min) was enough to obtain a
strong Gag signal, while limiting non-specific background staining.
As early as 20 hrs post-infection, specific Gag staining could be
readily detected. However, at earlier time points, the cytosolic Gag
signals were too low to be clearly distinguishable from background
fluorescence, and therefore no data were acquired before 20 h
postinfection.
Fluorescence microscopy and EM
For fluorescence microscopy, 24–48 hours post transfection/
infection, cells were labeled using the biarsenical method and either
fixed using 3.7% formaldehyde/PBS for 20 min or Jurkat T-cells
were added to the MDMs (in Opti-MEM I) for 2 hours, then fixed
with formaldehyde. The cells were then permeabilized with 0.1%
Triton-X100/PBS and incubated with 0.1 M glycine for 10 min at
room temperature to quench free aldehyde groups. The cells were
then blocked with 3% bovine serum albumin (BSA)/PBS, incubated
with either mouse monoclonal anti-CD63 (Santa Cruz Biotechnol-
ogy) or mouse monoclonal anti-CD81 (BD Pharmingen) for 1 hr at
room temperature, washed and incubated with Alexa-594 or 488-
conjugated anti-mouse IgG (Invitrogen) for 30 min at room
temperature. The cells were then washed and mounted with Aqua
Poly Mount (Polysciences Inc., Warrington, PA). For live cell
imaging, the labeled cells were imaged in a temperature-controlled
chamber (37uC/ 5% CO
2
) in Opti-MEM I. For both fixed and live-
cell microscopy, the cells were imaged using an Olympus 1X-71
inverted deconvolution microscope and analyzed with Delta Vision
software (Applied Precision Inc., Seattle, WA). To quantify the
degree of relative colocalization, we obtained the Pearson
correlation coefficient (R) values, which are standard measures of
colocalization [78]. The R values were calculated using the
softWORx colocalization module which generates a ‘‘colocalized’’
image from two channels. A scatter plot of the two intensities on a
pixel-by-pixel basis is then plotted and the R value is calculated by
dividing the covariances of each channel by the product of their
standard deviations. For EM, infected cells were fixed and processed
as previously described [15].
Metabolic labeling and radioimmunoprecipitation
analysis
Metabolic radiolabeling, preparation of cell and viral lysates,
and immunoprecipitation assays were performed as previously
described [15]. Briefly, transfected HeLa cells, or infected MDMs,
labeled with the biarsenical dyes or DMSO (control) were
metabolically labeled with [
35
S] Met/Cys for 2 hours, 24-48 hours
posttransfection/infection, and released virions were pelleted by
ultracentrifugation. Cell and virus lysates were immunoprecipitat-
ed with HIV immunoglobulin (HIV-Ig), obtained from NABI and
the National Heart Blood and Lung Institute through the NIH
AIDS Research and Reference Reagent Program. Immunopre-
cipitates were subjected to SDS-PAGE followed by fluorography.
Quantitative analysis of the bands visualized by radioimmunopre-
cipitation was performed using a Bio-rad phosphorimager.
Supporting Information
Figure S1 Colocalization of 29KE/31KE-TC Gag with CD63
in HeLa Cells. Cells transfected with pNL4-3/29KE/31KE-TC
HIV-1 Gag in Macrophages
PLoS Pathogens | www.plospathogens.org 12 2008 | Volume 4 | Issue 3 | e1000015
were labeled for 5 min with ReAsH (Gag) for 5 min at 37 uC,
washed, fixed in 3.7% formaldehyde and then stained with anti-
CD63 Ab (CD63). Lower panels show the overlay between the
Gag and CD63 staining. Scale bars = 15 mm.
Found at: doi:10.1371/journal.ppat.1000015.s001 (4.20 MB TIF)
Figure S2 Quantification of Gag/CD63 and Gag/CD81
Colocalization. MDMs infected with VSV-G-pseudotyped were
analyzed for Gag, CD63 and CD81 localization as described in
the Fig. 3 legend. Pearson coefficient (R) values were obtained for
a total of 75 cells and were plotted on a scale of 0 to 1 (x-axis). The
y-axis indicates the number of cells scored within each R-value
range.
Found at: doi:10.1371/journal.ppat.1000015.s002 (2.57 MB TIF)
Figure S3 Sample Images Used to Quantify the Localization of
Gag at the Synapse. Illustrates method used for data presented in
Fig. 6C (panel A) and Fig. 7C (panel B). The microscopy images
were opened in ImageJ and a plot profile was obtained. A column
average plot was generated, in which the x-axis represents the
horizontal distance through the image and the y-axis the vertically
averaged pixel intensity. The % Gag at the synapse was calculated
as: (pixel intensity of Gag signal at synapse)/(total intensity of Gag
in the cell+synapse)6100. N values indicate the number of cells
analyzed in each data set. Scale bar = 10 mm.
Found at: doi:10.1371/journal.ppat.1000015.s003 (3.67 MB TIF)
Video S1 Analysis of 29KE/31KE-TC Gag localization in
MDM in 3D. The z-series reconstructions were obtained by using
the Maximum Intensity Projection mode from the image
processing software OsiriX. The data correspond to the MDM
29KE/31KE-TC Gag panel in Fig. 2.
Found at: doi:10.1371/journal.ppat.1000015.s004 (0.18 MB
MOV)
Video S2 Analysis of Gag localization in 3D. The z-series
reconstructions were obtained by using the Maximum Intensity
Projection mode from the image processing software OsiriX. The
image in this figure corresponds to Fig. 5A.
Found at: doi:10.1371/journal.ppat.1000015.s005 (0.41 MB
MOV)
Video S3 Analysis of Gag localization in 3D. The z-series
reconstructions were obtained by using the Maximum Intensity
Projection mode from the image processing software OsiriX. The
image in this figure corresponds to Fig. 6A, bottom.
Found at: doi:10.1371/journal.ppat.1000015.s006 (0.38 MB
MOV)
Video S4 Analysis of Gag localization in 3D. The z-series
reconstructions were obtained by using the Maximum Intensity
Projection mode from the image processing software OsiriX. The
image in this figure corresponds to Fig. 6A, top.
Found at: doi:10.1371/journal.ppat.1000015.s007 (0.16 MB
MOV)
Video S5 Analysis of Gag localization in 3D. The z-series
reconstructions were obtained by using the Maximum Intensity
Projection mode from the image processing software OsiriX. The
image in this figure corresponds to Fig. 7A.
Found at: doi:10.1371/journal.ppat.1000015.s008 (0.30 MB
MOV)
Video S6 Analysis of Gag localization in 3D. The z-series
reconstructions were obtained by using the Maximum Intensity
Projection mode from the image processing software OsiriX. The
image in this figure corresponds to Fig. 8, top.
Found at: doi:10.1371/journal.ppat.1000015.s009 (0.13 MB
MOV)
Video S7 Movie of Gag movement to the MDM/MDM and
MDM/T-cell synapse. This experiment corresponds to the gallery
presented in Fig. 9A.
Found at: doi:10.1371/journal.ppat.1000015.s010 (8.86 MB
MOV)
Acknowledgments
We thank V. KewalRamani and members of the Freed lab for critical
review of the manuscript and helpful discussions. We thank K. Strebel for
plasmid Vpu-DEL-1. HIV-Ig and TZM-bl cells were obtained from NABI
and J. Kappes, respectively, through the NIH AIDS Research and
Reference Reagent Program.
Author Contributions
Conceived and designed the experiments: KG DEO EOF. Performed the
experiments: KG SDA LVC FS KN. Analyzed the data: KG AO DEO
EOF. Contributed reagents/materials/analysis tools: KG LVC AO DEO.
Wrote the paper: KG EOF.
References
1. Freed EO (1998) HIV-1 gag proteins: diverse functions in the virus life cycle.
Virology 251: 1–15.
2. Swanstrom R, Wills JW (1997) Synthesis, Assembly, and Processing of Viral
Proteins. In: Coffin JM, Hughes SH, Varmus HE, eds. Retroviruses. New York:
Cold Spring Harbor Laboratory Press. pp 263–334.
3. Adamson CS, Freed EO (2007) Human Immunodeficiency Type I Assembly,
Release and Maturation. Advances in Pharmacology S5: 347–387.
4. Bryant M, Ratner L (1990) Myristoylation-dependent replication and assembly
of human immunodefi ciency virus 1. Proc Natl Acad Sci U S A 87: 523–527.
5. Ono A, Ablan SD, Lockett SJ, Nagashima K, Freed EO (2004) Phosphatidy-
linositol (4,5) bisphosphate regulates HIV-1 Gag targeting to the plasma
membrane. Proc Natl Acad Sci U S A 101: 14889–14894.
6. Saad JS, Miller J, Tai J, Kim A, Ghanam RH, et al. (2006) Structural basis for
targeting HIV-1 Gag proteins to the plasma membrane for virus assembly. Proc
Natl Acad Sci U S A 103: 11364–11369.
7. Zhou W, Parent LJ, Wills JW, Resh MD (1994) Identification of a membrane-
binding domain within the amino-terminal region of human immunodeficiency
virus type 1 Gag protein which interacts with acidic phospholipids. J Virol 68:
2556–2569.
8. von Schwedler UK, Stray KM, Garrus JE, Sundquist WI (2003) Functional
surfaces of the human immunodeficiency virus type 1 capsid protein. J Virol 77:
5439–5450.
9. Rein A, Henderson LE, Levin JG (1998) Nucleic-acid-chaperone activity of
retrovir al nucleocapsid proteins: significance for viral replication. Trends
Biochem Sci 23: 297–301.
10. Bieniasz PD (2006) Late budding domains and host proteins in enveloped virus
release. Virology 344: 55–63.
11. Demirov DG, Freed EO (2004) Retrovirus budding. Virus Res 106: 87–102.
12. Morita E, Sundquist WI (2004) Retrovirus budding. Annu Rev Cell Dev Biol 20:
395–425.
13. Facke M, Janetzko A, Shoeman RL, Krausslich HG (1993) A large deletion in
the matrix domain of the human immunodeficiency virus gag gene redirects
virus particle assembly from the plasma membrane to the endoplasmic
reticulum. J Virol 67: 4972–4980.
14. Gallina A, Mantoan G, Rindi G, Milanesi G (1994) Influenc e of MA
internal sequences, but not of the myristylated N-terminus sequence, on the
budding site of HIV-1 Gag protein. Biochem Biophys Res Commun 204:
1031–1038.
15. Freed EO, Orenstein JM, Buckler-White AJ, Martin MA (1994) Single amino
acid changes in the human immunodeficiency virus type 1 matrix protein block
virus particle production. J Virol 68: 5311–5320.
16. Ono A, Orenstein JM, Freed EO (2000) Role of the Gag matrix domain in
targeting human immunodeficien cy virus type 1 assembly. J Virol 74:
2855–2866.
17. Yuan X, Yu X, Lee TH, Essex M (1993) Mutations in the N-terminal region of
human immunodeficiency virus type 1 matrix protein block intracellular
transport of the Gag precursor. J Virol 67: 6387–6394.
18. Ono A, Freed EO (2004) Cell-type-dependent targeting of human immunode-
ficiency virus type 1 assembly to the plasma membrane and the multivesicular
body. J Virol 78: 1552–1563.
HIV-1 Gag in Macrophages
PLoS Pathogens | www.plospathogens.org 13 2008 | Volume 4 | Issue 3 | e1000015
19. Booth AM, Fang Y, Fallon JK, Yang JM, Hildreth JE, et al. (2006) Exosomes
and HIV Gag bud from endosome-like domains of the T cell plasma membrane.
J Cell Biol 172: 923–935.
20. Gould SJ, Booth AM, Hildreth JE (2003) The Trojan exosome hypothesis. Proc
Natl Acad Sci U S A 100: 10592–10597.
21. Grigorov B, Arcanger F, Roingeard P, Darlix JL, Muriaux D (2006) Assembly of
infectious HIV-1 in human epithelial and T-lymphoblastic cell lines. J Mol Biol
359: 848–862.
22. Nydegger S, Foti M, Derdowski A, Spearman P, Thali M (2003) HIV-1 egress is
gated through late endosomal membranes. Traffic 4: 902–910.
23. Perlman M, Resh MD (2006) Identification of an intracellular trafficking and
assembly pathway for HIV-1 gag. Traffic 7: 731–745.
24. Sherer NM, Lehmann MJ, Jimenez-Soto LF, Ingmundson A, Horner SM, et al.
(2003) Visualization of retroviral replication in living cells reveals budding into
multivesicular bodies. Traffic 4: 785–801.
25. Finzi A, Orthwein A, Mercier J, Cohen EA (2007) Productive human
immunodeficiency virus type 1 assembly takes place at the plasma membrane.
J Virol 81: 7476–74 90.
26. Harila K, Prior I, Sjoberg M, Salminen A, Hinkula J, et al. (2006) Vpu and
Tsg101 regulate intracellular targeting of the human immunode ficiency virus
type 1 core protein precursor Pr55gag. J Virol 80: 3765–3772.
27. Jouvenet N, Neil SJ, Bess C, Johnson MC, Virgen CA, et al. (2006) Plasma
membrane is the site of productive HIV-1 particle assembly. PLoS Biol 4: e435.
28. Neil SJ, Eastman SW, Jouvenet N, Bieniasz PD (2006) HIV-1 Vpu promotes
release and prevents endocytosis of nascent retrovirus particles from the plasma
membrane. PLoS Pathog 2: e39.
29. Rudner L, Nydegger S, Coren LV, Nagashima K, Thali M, et al. (2005)
Dynamic fluorescent imaging of human immunodeficiency virus type 1 gag in
live cells by biarsenical labeling. J Virol 79: 4055–4065.
30. Joshi A, Freed EO (2007) HIV-1 Gag trafficking. Future HIV Ther 1: 427–438.
31. Gendelman HE, Orenstein JM, Martin MA, Ferrua C, Mitra R, et al. (1988)
Efficient isolation and propagation of human immunodeficiency virus on
recombinant colony-stimulating factor 1-treated monocytes. J Exp Med 167:
1428–1441.
32. Orenstein JM, Jannotta F (1988) Human immunodeficiency virus and
papovavirus infections in acquired immunodeficiency syndrome: an ultrastruc-
tural study of three cases. Hum Pathol 19: 350–361.
33. Pelchen-Matthews A, Kramer B, Marsh M (2003) Infectious HIV-1 assembles in
late endosomes in primary macrophages. J Cell Biol 162: 443–455.
34. Raposo G, Moore M, Innes D, Leijendekker R, Leigh-Brown A, et al. (2002)
Human macrophages accumulate HIV-1 particles in MHC II compartments.
Traffic 3: 718–729.
35. Chertova E, Chertov O, Coren LV, Roser JD, Trubey CM, et al. (2 006)
Proteomic and biochemical analysis of purified human immunodeficiency virus
type 1 produced from infected monocyte-derived macrophages. J Virol 80:
9039–9052.
36. Nguyen DG, Booth A, Gould SJ, Hildreth JE (2003) Evidence that HIV budding
in primary macrophages occurs through the exosome release pathway. J Biol
Chem 278: 52347–52354.
37. Deneka M, Pelchen-Matthews A, Byland R, Ruiz-Mateos E, Marsh M (2007) In
macrophages, HIV-1 assembles into an intracellular plasma membrane domain
containing the tetraspanins CD81, CD9, and CD53. J Cell Biol 177: 329–341.
38. Welsch S, Keppler OT, Habermann A, Allespach I, Krijnse-Locker J, et al.
(2007) HIV-1 buds predominantly at the plasma membrane of primary human
macrophages. PLoS Pathog 3: e36.
39. Koenig S, Gendelman HE, Orenstein JM, Dal Canto MC, Pezeshkpour GH, et
al. (1986) Detection of AIDS virus in macrophages in brain tissue from AIDS
patients with encephalopathy. Science 233: 1089–1093.
40. Adams SR, Campbell RE, Gross LA, Martin BR, Walkup GK, et al. (2002) New
biarsenical ligands and tetracysteine motifs for protei n labeling in vitro and in
vivo: synthesis and biological applications. J Am Chem Soc 124: 6063–6076.
41. Gaietta G, Deerinck TJ, Adams SR, Bouwer J, Tour O, et al. (2002) Multicolor
and electron microscopic imaging of connexin trafficking. Science 296: 503–507.
42. Griffin BA, Adams SR, Tsien RY (1998) Specific covalent labeling of
recombinant protein molecules inside live cells. Science 281: 269–272.
43. Yu X, Yu QC, Lee TH, Essex M (1992) The C terminus of human
immunodeficiency virus type 1 matrix protein is involved in early steps of the
virus life cycle. J Virol 66: 5667–5670.
44. Muller B, Daecke J, Fackler OT, Dittmar MT, Zentgraf H, et al. (2004)
Construction and characterization of a fluorescently labeled infectious human
immunodeficiency virus type 1 derivative. J Virol 78: 10803–10813.
45. Freed EO, Englund G, Martin MA (1995) Role of the basic domain of human
immunodeficiency virus type 1 matrix in macrophage infection. J Virol 69:
3949–3954.
46. Freed EO, Martin MA (1994) HIV-1 infection of non-dividing cells. Nature 369:
107–108.
47. Platt EJ, Wehrly K, Kuhmann SE, Chesebro B, Kabat D (1998) Effects of
CCR5 and CD4 cell surface concentrations on infections by macrophagetropic
isolates of human immunodeficiency virus type 1. J Virol 72: 2855–2864.
48. Ono A, Demir ov D, F reed EO (2000) Rela tionship between human
immunodeficiency virus type 1 Gag multimerization and membrane binding.
J Virol 74: 5142–5150.
49. Joshi A, Nagashima K, Freed EO (2006) Mutation of dileucine-like motifs in the
human immunodeficiency virus type 1 capsid disrupts virus assembly, gag-gag
interactions, gag-membrane binding, and virion matu ration. J Virol 80:
7939–7951.
50. Nydegger S, Khurana S, Krementsov DN, Foti M, Thali M (2006) Mapping of
tetraspanin-enriched microdomains that can function as gateways for HIV-1.
J Cell Biol 173: 795–807.
51. Bromley SK, Burack WR, Johnson KG, Somersalo K, Sims TN, et al. (2001)
The immunological synapse. Annu Rev Immunol 19: 375–396.
52. Grakoui A, Bromley SK, Sumen C, Davis MM, Shaw AS, et al. (1999) The
immunological synapse: a molecular machine controlling T cell activation.
Science 285: 221–227.
53. Monks CR, Freiberg BA, Kupfer H, Sciaky N, Kupfer A (1998) Three-
dimensional segregation of supramolecular activation clusters in T cells. Nature
395: 82–86.
54. Arrighi JF, Pion M, Garcia E, Escola JM, van Kooyk Y, et al. (2004) DC-SIGN-
mediated infectious synapse formation enhances X4 HIV-1 transmission from
dendritic cells to T cells. J Exp Med 200: 1279–1288.
55. Jolly C, Sattentau QJ (2004) Retroviral spread by induction of virological
synapses. Traffic 5: 643–650.
56. McDonald D, Wu L, Bohks SM, KewalRamani VN, Unutmaz D, et al. (2003)
Recruitment of HIV and its receptors to dendritic cell-T cell junctions. Science
300: 1295–1297.
57. Garcia E, Pion M, Pelchen-Matthews A, Collinson L, Arrighi JF, et al. (2005)
HIV-1 trafficking to the dendritic cell-T-cell infectious synapse uses a pathway of
tetraspanin sorting to the immunological synapse. Traffic 6: 488–501.
58. Wu L, KewalRamani VN (2006) Dendritic-cell interactions with HIV: infection
and viral dissemination. Nat Rev Immunol 6: 859–868.
59. Jolly C, Kashefi K, Hollinshead M, Sattentau QJ (2004) HIV-1 cell to cell
transfer across an Env-induced, actin-dependent synapse. J Exp Med 199:
283–293.
60. Sherer NM, Lehmann MJ, Jimenez-Soto LF, Horensavitz C, Pypaert M, et al.
(2007) Retroviruses can establish filopodial bridges for efficient cell-to-cell
transmission. Nat Cell Biol 9: 310–315.
61. Nagai T, Ibata K, Park ES, Kubota M, Mikoshiba K, et al. (2002) A variant of
yellow fluorescent protein with fast and efficient maturation for cell-biological
applications. Nat Biotechnol 20: 87–90.
62. Remington SJ (2006) Fluorescent proteins: maturation, photochemistry and
photophysics. Curr Opin Struct Biol 16: 714–721.
63. Sharova N, Swingler C, Sharkey M, Stevenson M (2005) Macrophages archive
HIV-1 virions for dissemination in trans. Embo J 24: 2481–2489.
64. Carr JM, Hocking H, Li P, Burrell CJ (1999) Rapid and efficient cell-to-cell
transmission of human immunodeficiency virus infection from monocyte-derived
macrophages to peripheral blood lymphocytes. Virology 265: 319–329.
65. Turville SG, Santos JJ, Frank I, Cameron PU, Wilkinson J, et al. (2004)
Immunodeficiency virus uptake, turnover, and 2-phase transfer in human
dendritic cells. Blood 103: 2170–2179.
66. Wang JH, Janas AM, Olson WJ, Wu L (2007) Functionally distinct transmission
of human immunodeficiency virus type 1 mediated by immature and mature
dendritic cells. J Virol 81: 8933–8943.
67. Jouve M, Sol-Foulon N, Watson S, Schwartz O, Benaroch P (2007) HIV-1 buds
and accumulates in ‘‘nonacidic’’ endosomes of macrophages. Cell Host Microbe
2: 85–95.
68. Jolly C, Sattentau QJ (2007) Human immunodeficiency virus type 1 assembly,
budding, and cell-cell spread in T cells take place in tetraspanin-enriched plasma
membrane domains. J Virol 81: 7873–7884.
69. Sylwester A, Wessels D, Anderson SA, Warren RQ, Shutt DC, et al. (1993)
HIV-induced syncytia of a T cell line form single giant pseudopods and are
motile. J Cell Sci 106 ( Pt 3): 941–953.
70. Adachi A, Gendelman HE, Koenig S, Folks T, Willey R, et al. (1986) Production
of acquired immunodeficiency syndrome-associated retrovirus in human and
nonhuman cells transfected with an infectious molecular clone. J Virol 59:
284–291.
71. Horton RM, Cai ZL, Ho SN, Pease LR (1990) Gene splicing by overlap
extension: tailor-made genes using the polymerase chain reaction. Biotechniques
8: 528–535.
72. Freed EO, Delwart EL, Buchschacher GL Jr, Panganiban AT (1992) A mutation
in the human immunodeficiency virus type 1 transmembrane glycoprotein gp41
dominantly interferes with fusion and infectivity. Proc Natl Acad Sci U S A 89:
70–74.
73. Klimkait T, Strebel K, Hoggan MD, Martin MA, Orenstein JM (1990) The
human immunodeficiency virus type 1-specific protein vpu is required for
efficient virus maturation and release. J Virol 64: 621–629.
74. Ono A, Freed EO (2001) Plasma membrane rafts play a critical role in HIV-1
assembly and release. Proc Natl Acad Sci U S A 98: 13925–13930.
75. Yee JK, Friedmann T, Burn s JC (1994) Generation of high-titer pseudotyped
retroviral vectors with very broad host range. Methods Cell Biol 43 Pt A:
99–112.
76. Willey RL, Bonifaci no JS, Potts BJ, Martin MA, Klausner R D (1988)
Biosynthesis, cleavage, and degradation of the human immunodeficiency virus
1 envelope glycoprotein gp160. Proc Natl Acad Sci U S A 85: 9580–9584.
77. Kiernan RE, Ono A, Englund G, Freed EO (1998) Role of matrix in an early
postentry step in the human immunodeficiency virus type 1 life cycle. J Virol 72:
4116–4126.
78. Manders EM, Verbeek FJ, Aten JA (1992) Measurement of colocalization of
objects in dual-colour confocal images. J Micro 169: 375–382.
HIV-1 Gag in Macrophages
PLoS Pathogens | www.plospathogens.org 14 2008 | Volume 4 | Issue 3 | e1000015
... replication 5,9,13,14,[34][35][36][37][38] . As such, we next determined if C99 or the C99-7KA mutant affected Gag colocalization with CD63-positive vesicles. ...
... Indeed, beyond these early events, as infection progresses the majority of Gag accumulates in modified vesicles that contain markers of LE's or MVB's, in part through continued internalization from the cell surface but largely by direct budding of Gag into MVBs 11,39-43 . Studies from several groups have shown that at these later stages of infection, Gag buds into MVB's using the vesicular VPS machinery and that this is a major productive pathway in macrophages [11][12][13][14][15]35,36,[38][39][40][41][42][43][44][45][46][47][48] . The MVB subsets that support virus replication are characterized by low levels of LAMP1 and tetraspanins such CD53, but high levels of CD63 9,13,14,34-38 . ...
Article
Full-text available
HIV-1 replication in macrophages and microglia involves intracellular assembly and budding into modified subsets of multivesicular bodies (MVBs), which support both viral persistence and spread. However, the cellular factors that regulate HIV-1’s vesicular replication remain poorly understood. Recently, amyloid precursor protein (APP) was identified as an inhibitor of HIV-1 replication in macrophages and microglia via an unknown mechanism. Here, we show that entry of HIV-1 Gag into MVBs is blocked by the amyloidogenic C-terminal fragment of APP, “C99”, but not by the non-amyloidogenic product, “C83”. To counter this, Gag promotes multi-site ubiquitination of C99 which controls both exocytic sorting of MVBs and further processing of C99 into toxic amyloids. Processing of C99, entry of Gag into MVBs and release of infectious virus could be suppressed by expressing ubiquitination-defective C99 or by γ-secretase inhibitor treatment, suggesting that APP’s amyloidogenic pathway functions to sense and suppress HIV-1 replication in macrophages and microglia.
... Cell-to-cell transmission may be more efficient than the transmission of cell-free virions (Sato et al. 1992;Phillips 1994). Cell-to-cell transmission occurs at the VS when tight clefts are formed between HIV-1-infected effector cells and target cells (Sato et al. 1992;Gousset et al. 2008;Real et al. 2018). Following VS formation, viral budding polarizes toward the target cell. ...
... Following VS formation, viral budding polarizes toward the target cell. Budding viruses, viral RNA, and virus-related proteins are likely transmitted to target cells through this cleft (Gousset et al. 2008;Real et al. 2018). The movement of HIV-1 across adhesive junctions might protect the virus from humoral immune responses, antiretroviral drugs, neutralizing antibodies, and complements (Abela et al. 2012;Kulpa et al. 2013). ...
Article
Full-text available
Galectin-3 (GAL3) is a β-galactoside-binding lectin expressed in CD4 T cells infected with human immunodeficiency virus-1 (HIV-1). GAL3 promotes HIV-1 budding by associating with ALIX and Gag p6. GAL3 has been shown to localize in membrane lipid rafts in dendritic cells and positively regulate cell migration. HIV-1 spreads between T cells by forming supramolecular structures (virological synapses [VSs]), whose integrity depends on lipid rafts. Here, we addressed the potential role of GAL3 in cell-to-cell transmission of HIV-1 in CD4 T cells. GAL3 expressed in donor cells was more important for facilitating HIV-1 cell-to-cell transfer than GAL3 expressed in target cells. GAL3 was found to be co-transferred with Gag from HIV-1-positive donor to HIV-1-negative target T cells. HIV-1 infection induced translocation of GAL3 together with Gag to the cell–cell interfaces and colocalize with GM1, where GAL3 facilitated VS formation. GAL3 regulated the coordinated transfer of Gag and flotillin-1 into plasma membrane fractions. Finally, depletion of GAL3 reduced the cholesterol levels in membrane lipid rafts in CD4 T cells. These findings provide evidence that endogenous GAL3 stimulates lipid raft components and facilitates intercellular HIV-1 transfer among CD4 T cells, offering another pathway by which GAL3 regulates HIV-1 infection. These findings may inform the treatment of HIV-1 infection based on targeting GAL3 to modulate lipid rafts.
... However, in our cultures, direct viral transmission from one cell to another does not occur due to the absence of the viral envelope. VCCs are often observed in proximity to the viral synapse that forms between an infected and a target cell (Duncan et al., 2014;Gousset et al., 2008), where they potentially participate in direct viral transmission. Interestingly, direct transmission of HIV-1 from macrophages to CD4 + T cells is inhibited by jasplakinolide (Duncan et al., 2014), suggesting that the same actin-dependent mechanisms that regulate VCC trafficking for viral release also impact its redistribution towards viral synapses. ...
Article
A feature of HIV-1 replication in macrophages is that viral assembly occurs at the limiting membrane of a compartment often named VCC (virus-containing compartments). Assembled virions accumulate in the lumen of the VCC, from where they can be released into the extracellular medium, via mechanisms that remain poorly described. Here we show that the actin cytoskeleton contributes to this process by combining pharmacological and mechanical perturbations with imaging and biochemical analysis. We found that jasplakinolide inhibited HIV-1 release from macrophages and led to scattering of the compartment. Concomitantly, both the integrin CD18 and the phosphorylated form of PYK2 were displaced away from the VCC. Inhibition of PYK2 activity promoted retention of viral particles in VCC that lost their connections to the surface. Finally, in infected macrophages undergoing frustrated phagocytosis, VCC rapidly trafficked to the basal membrane and released their viral content, in a manner dependent on their association with the actin cytoskeleton. These results highlight that VCC trafficking and virus release are intimately linked to the reorganization of the macrophage actin cytoskeleton that can be modulated by external physical cues.
... Enfin, l'acheminement du complexe Gag-ARNv vers les rafts ou les endosomes tardifs sera assuré par la matrice de Gag qui, suite à son groupement myristate en N-ter associé à des résidus basiques permet l'ancrage de Gag dans les lipides membranaires de la cellule (159)(160)(161)(162) (182). Une mutation dans ce domaine affect la traduction de différentes protéines expliquant l'importance de ABCE1 pour la viabilité des cellules (183). ...
Thesis
A la fin du cycle viral, la polyprotéine Gag du VIH-1 polymérise afin de former la particule virale et recrute les partenaires viraux et cellulaires permettant le bourgeonnement de particules infectieuses. De plus, par son activité de chaperonne des acides nucléiques, elle induit la dimérisation et l’encapsidation de l’ARN viral. Plusieurs études in vitro ont montré que Gag possédait une faible activité de chaperonne. Nous avons donc cherché à identifier un partenaire cellulaire qui, en étant recruté par Gag, serait capable d’en améliorer l’activité. Parmi les partenaires identifiés, nous nous sommes intéressés à la protéine ribosomale RPL7. Nous avons caractérisé le complexe Gag-RPL7 et montré que le domaine NCp7 de Gag interagissait avec les parties N et C terminales de RPL7. En se basant sur un test in vitro consistant à suivre l’hybridation de cTAR avec dTAR, nous avons montré que la RPL7 possédait une activité chaperonne des acides nucléiques supérieure à celle de Gag et que l’activité du complexe Gag-RPL7 était supérieur à celle de chaque protéine. Enfin, nous avons déchiffré le mécanisme d’hybridation de cTAR/dTAR et montré que la RPL7, Gag ou Gag-RPL7, formaient l’hybride cTAR/dTAR suivant différentes voies. Tous ces résultats nous ont permis de proposer un modèle dans lequel, une fois exprimée, Gag serait capable de recruter la protéine RPL7 pour améliorer son activité de chaperonne.
... HIV-1 j retrovirus j Gag j matrix j envelope During the late phase of HIV type 1 (HIV-1) infection cycle, the virally encoded Gag polyproteins are targeted to the inner leaflet of the plasma membrane (PM) for assembly of the immature particle and subsequent budding and release (1)(2)(3)(4)(5)(6)(7)(8)(9). Concomitant or subsequent to particle budding, the Gag protein is cleaved by the viral protease into matrix (MA), capsid (CA), nucleocapsid, and short peptides (SP1, SP2, and P6), leading to formation of mature particles (reviewed in refs. 1 and 2). ...
Article
Full-text available
Significance The assembly of immature HIV-1 particles is initiated by targeting of the Gag polyproteins to the plasma membrane (PM). Gag binding to the PM is mediated by the N-terminally myristoylated matrix (myrMA) domain. Formation of a Gag lattice on the PM is obligatory for the assembly of immature HIV-1 and envelope (Env) incorporation. The structure of the myrMA lattice presented here provided insights on the molecular factors that stabilize the lattice and hence favor Env incorporation. Our data support a mechanism for Gag binding to the PM during the assembly of immature particles and upon maturation. These findings advance our understanding of a critical step in HIV-1 assembly.
... Single-cell video microscopy has been used to study several independent features or stages of the HIV-1 viral life cycle, including, but not limited to: viral entry [16,17], postentry events [18][19][20][21][22], transcription [23,24], US vRNA nuclear export [25,26], translation [27,28], vRNA and Gag transport in the cytoplasm [29][30][31][32][33], virus particle assembly [34][35][36][37][38][39][40], and the cell-to-cell transfer of virions [41][42][43][44][45]. To date, however, many of these studies have been focused on single viral activities that occur over relatively short time intervals (e.g., seconds to minutes). ...
Article
Full-text available
Single-cell imaging has emerged as a powerful means to study viral replication dynamics and identify sites of virus–host interactions. Multivariate aspects of viral replication cycles yield challenges inherent to handling large, complex imaging datasets. Herein, we describe the design and implementation of an automated, imaging-based strategy, “Human Immunodeficiency Virus Red-Green-Blue” (HIV RGB), for deriving comprehensive single-cell measurements of HIV-1 unspliced (US) RNA nuclear export, translation, and bulk changes to viral RNA and protein (HIV-1 Rev and Gag) subcellular distribution over time. Differentially tagged fluorescent viral RNA and protein species are recorded using multicolor long-term (>24 h) time-lapse video microscopy, followed by image processing using a new open-source computational imaging workflow dubbed “Nuclear Ring Segmentation Analysis and Tracking” (NR-SAT) based on ImageJ plugins that have been integrated into the Konstanz Information Miner (KNIME) analytics platform. We describe a typical HIV RGB experimental setup, detail the image acquisition and NR-SAT workflow accompanied by a step-by-step tutorial, and demonstrate a use case wherein we test the effects of perturbing subcellular localization of the Rev protein, which is essential for viral US RNA nuclear export, on the kinetics of HIV-1 late-stage gene regulation. Collectively, HIV RGB represents a powerful platform for single-cell studies of HIV-1 post-transcriptional RNA regulation. Moreover, we discuss how similar NR-SAT-based design principles and open-source tools might be readily adapted to study a broad range of dynamic viral or cellular processes.
Preprint
Full-text available
HIV-1 infects CD4+ T cells and macrophages. However, replication of HIV-1 in these cell types is highly variable and may depend on the use of CCR5 as a co-receptor. In addition, there is internal accumulation of infectious HIV-1 in so-called virus-containing compartments of macrophages (VCCs). VCCs are thought to represent a persistent viral reservoir that is shielded from the antiviral immune response. To date, VCC formation has only been studied in lab-adapted HIV-1 and it is unknown whether VCCs play a role in the replication of primary HIV-1 strains. Furthermore, although macrophages transmit HIV-1 from VCCs to CD4+ T cells, it is unknown whether T cells have an impact on VCC formation. We analyzed the ability of primary and lab-adapted HIV-1 to replicate in macrophages, the effect of coculture with non-infected CD4+ T cells and the extent of VCC formation. Although differentially, all HIV-1 strains replicated in CD4+ T cells, whereas only lab-adapted HIV-1 replicated in macrophages. Strikingly, replication of patient-derived HIV-1 in macrophages was enhanced by coculture with non-infected CD4+ T cells and correlated with VCC formation. In conclusion, non-infected CD4+ T cells facilitate the replication of primary HIV-1 strains in macrophages and the formation of VCCs appears to be a proxy for this phenotype. Our study suggests an essential role for VCCs in the replication of patient-derived HIV-1 in macrophages, which is fueled by non-infected CD4+ T cells. Furthermore, our findings call for strategies to specifically disrupt VCC formation in order to eliminate the HIV-1 reservoir in macrophages. IMPORTANCE Here we focus on the intimate interplay between HIV-1 infected macrophages and CD4+ T cells. Specifically, we analyzed whether primary HIV-1 strains induce virus-containing compartments (VCCs) within macrophages, which are thought to serve as viral sanctuaries and macrophage reservoirs. Notably, primary HIV-1 strains were unable to replicate in macrophages and induce VCCs unless they were cocultured with non-infected CD4+ T cells, leading to increased VCC formation and viral replication. This suggests an essential role for non-infected CD4+ T cells in facilitating primary HIV-1 replication in macrophages. Our data highlight the importance of not only targeting the latent HIV-1 T-cell reservoir, but also targeting VCC formation in macrophages to achieve the ultimate goal of functional HIV-1 cure.
Article
Despite the success of combined antiretroviral therapy in controlling viral load and reducing the risk of human immunodeficiency virus (HIV) transmission, an estimated 1.5 million new infections occurred worldwide in 2021. These new infections are mainly the result of sexual intercourse and thus involve cells present on the genital mucosa, such as dendritic cells (DCs), macrophages (Mø) and CD4+ T lymphocytes. Understanding the mechanisms by which HIV interacts with these cells and how HIV exploits these interactions to establish infection in a new human host is critical to the development of strategies to prevent and control HIV transmission. In this review, we explore how HIV has evolved to manipulate some of the physiological roles of these cells, thereby gaining access to strategic cellular niches that are critical for the spread and pathogenesis of HIV infection. The interaction of HIV with DCs, Mø and CD4+ T lymphocytes, and the role of the intercellular transfer of viral particles through the establishment of the infectious or virological synapses, but also through membrane protrusions such as filopodia and tunnelling nanotubes (TNTs), and cell fusion or cell engulfment processes are presented and discussed.
Article
Assembly of human T-cell leukemia virus type 1 (HTLV-1) particles is initiated by the trafficking of virally encoded Gag polyproteins to the inner leaflet of the plasma membrane (PM). Gag–PM interactions are mediated by the matrix (MA) domain, which contains a myristoyl group (myr) and a basic patch formed by lysine and arginine residues. For many retroviruses, Gag–PM interactions are mediated by phosphatidylinositol 4,5-bisphosphate [PI(4,5)P2]; however, previous studies suggested that HTLV-1 Gag–PM interactions and therefore virus assembly are less dependent on PI(4,5)P2. We have recently shown that PI(4,5)P2 binds directly to HTLV-1 unmyristoylated MA [myr(–)MA] and that myr(–)MA binding to membranes is significantly enhanced by inclusion of phosphatidylserine (PS) and PI(4,5)P2. Herein, we employed structural, biophysical, biochemical, and cell-based assays to identify residues involved in MA–membrane interactions. Our data revealed that the lysine-rich motif (Lys47, Lys48, and Lys51) constitutes the primary PI(4,5)P2–binding site. Furthermore, we show that arginine residues 3, 7, 14 and 17 located in the unstructured N-terminus are essential for MA binding to membranes containing PS and/or PI(4,5)P2. Both the lysine and arginine residues were confirmed to severely attenuate virus-like particle production, but only the lysine residues could be clearly correlated with reduced plasma membrane binding. These results support a mechanism by which HTLV-1 Gag targeting to the PM is mediated by a trio engagement of the myr group, Arg-rich and Lys-rich motifs. These findings advance our understanding of a key step in retroviral particle assembly.
Thesis
Les cellules dendritiques (DC) constituent une population hétérogène qui comporte les pDC, les cDC1, les cDC2 et les Axl+ DC récemment décrites dans le sang. Cependant, il reste à évaluer si les Axl+ DC possèdent des fonctions spécifiques après exposition à un pathogène. Elles expriment constitutivement Siglec-1, impliqué dans la capture du VIH-1, ce qui nous a incité à étudier leurs relations avec le VIH-1. Nous avons montré que seules les Axl+ DC étaient susceptibles à l’infection de manière dépendante de Siglec-1. Elles produisent des virus infectieux qui s’accumulent dans des compartiments semblables à ceux observés chez les macrophages. De plus, après activation, les Axl+ DC sont capables de capturer le VIH-1 via Siglec-1 et de le transmettre aux lymphocytes T CD4+, sans réplication virale. Les Axl+ DC possèdent également des spécificités fonctionnelles dans la détection et la réponse au VIH-1. Nous avons montré qu’elles étaient plus efficaces que les cDC2 à induire une réponse interféron et une réponse inflammatoire. Après exposition au VIH-1, ces réponses sont engagées dans des Axl+ DC différentes. En revanche, les Axl+ DC qui produisent de nouveaux virus présentent une activation simultanée de ces deux réponses. Nos résultats suggèrent que cette diversité fonctionnelle serait due à des mécanismes de détection du virus différents impliquant soit les TLR endosomaux, soit la voie STING. Nos travaux décrivent pour la première fois une fonction des Axl+ DC dans un contexte infectieux. Ils illustrent la complexité des relations entre les sous-types de DC et le VIH-1 dont l’analyse est nécessaire pour mieux comprendre les processus d’activation de la réponse immunitaire.
Article
Full-text available
The adaptive immune response is initiated by the interaction of T cell antigen receptors with major histocompatibility complex molecule-peptide complexes in the nanometer scale gap between a T cell and an antigen-presenting cell, referred to as an immunological synapse. In this review we focus on the concept of immunological synapse formation as it relates to membrane structure, T cell polarity, signaling pathways, and the antigen-presenting cell. Membrane domains provide an organizational principle for compartmentalization within the immunological synapse. T cell polarization by chemokines increases T cell sensitivity to antigen. The current model is that signaling and formation of the immunological synapse are tightly interwoven in mature T cells. We also extend this model to natural killer cell activation, where the inhibitory NK synapse provides a striking example in which inhibition of signaling leaves the synapse in its nascent, inverted state. The APC may also play an active role in immunological synapse formation, particularly for activation of naive T cells.
Article
Full-text available
It has been proposed that changes in cell surface concentrations of coreceptors may control infections by human immunodeficiency virus type 1 (HIV-1), but the mechanisms of coreceptor function and the concentration dependencies of their activities are unknown. To study these issues and to generate stable clones of adherent cells able to efficiently titer diverse isolates of HIV-1, we generated two panels of HeLa-CD4/CCR5 cells in which individual clones express either large or small quantities of CD4 and distinct amounts of CCR5. The panels were made by transducing parental HeLa-CD4 cells with the retroviral vector SFF-CCR5. Derivative clones expressed a wide range of CCR5 quantities which were between 7.0 x 10(2) and 1.3 x 10(5) molecules/cell as measured by binding antibodies specific for CCR5 and the chemokine [125I]MIP1beta. CCR5 was mobile in the membranes, as indicated by antibody-induced patching. In cells with a large amount of CD4, an unexpectedly low trace of CCR5 (between 7 x 10(2) and 2.0 x 10(3) molecules/cell) was sufficient for maximal susceptibility to all tested HIV-1, including primary patient macrophagetropic and T-cell-tropic isolates. Indeed, the titers as indicated by immunoperoxidase staining of infected foci were as high as the tissue culture infectious doses measured in human peripheral blood mononuclear cells. In contrast, cells with a small amount of CD4 required a much larger quantity of CCR5 for maximal infection by macrophagetropic HIV-1 (ca. 1.0 x 10(4) to 2.0 x 10(4) molecules/cell). Cells that expressed low and high amounts of CD4 were infected with equal efficiencies when CCR5 concentrations were above threshold levels for maximal infection. Our results suggest that the concentrations of CD4 and CCR5 required for efficient infections by macrophagetropic HIV-1 are interdependent and that the requirements for each are increased when the other component is present in a limiting amount. We conclude that CD4 and CCR5 directly or indirectly interact in a concentration-dependent manner within a pathway that is essential for infection by macrophagetropic HIV-1. In addition, our results suggest that multivalent virus-receptor bonds and diffusion in the membrane contribute to HIV-1 infections.
Article
Deneka et al. 2007. J. Cell Biol. doi:10.1083/jcb.200609050 [OpenUrl][1][Abstract/FREE Full Text][2] [1]: {openurl}?query=rft_id%253Dinfo%253Adoi%252F10.1083%252Fjcb.200609050%26rft_id%253Dinfo%253Apmid%252F17438075%26rft.genre%253Darticle%26rft_val_fmt%253Dinfo%253Aofi%252Ffmt%253Akev%253Amtx%
Article
We propose that retroviruses exploit a cell-encoded pathway of intercellular vesicle traffic, exosome exchange, for both the biogenesis of retroviral particles and a low-efficiency but mechanistically important mode of infection. This Trojan exosome hypothesis reconciles current paradigms of retrovirus-directed transmission with the unique lipid composition of retroviral particles, the host cell proteins present in retroviral particles, the complex cell biology of retroviral release, and the ability of retroviruses to infect cells independently of Envelope protein-receptor interactions. An exosomal origin also predicts that retroviruses pose an unsolvable paradox for adaptive immune responses, that retroviral antigen vaccines are unlikely to provide prophylactic protection, and that alloimmunity is a central component of antiretroviral immunity. Finally, the Trojan exosome hypothesis has important implications for the fight against HIV and AIDS, including how to develop new antiretroviral therapies, assess the risk of retroviral infection, and generate effective antiretroviral vaccines.
Article
Recombinant proteins containing four cysteines at the i,i + 1, i + 4, and i + 5 positions of an α helix were fluorescently labeled in living cells by extracellular administration of 4′,5′-bis(1,3,2-dithioarsolan-2-yl)fluorescein. This designed small ligand is membrane-permeant and nonfluorescent until it binds with high affinity and specificity to the tetracysteine domain. Such in situ labeling adds much less mass than does green fluorescent protein and offers greater versatility in attachment sites as well as potential spectroscopic and chemical properties. This system provides a recipe for slightly modifying a target protein so that it can be singled out from the many other proteins inside live cells and fluorescently stained by small nonfluorescent dye molecules added from outside the cells.
Article
We recently introduced a method (Griffin, B. A.; Adams, S. R.; Tsien, R. Y. Science 1998, 281, 269-272 and Griffin, B. A.; Adams, S. R.; Jones, J.; Tsien, R. Y. Methods Enzymol. 2000, 327, 565-578) for site-specific fluorescent labeling of recombinant proteins in living cells. The sequence Cys-Cys-Xaa-Xaa-Cys-Cys, where Xaa is an noncysteine amino acid, is genetically fused to or inserted within the protein, where it can be specifically recognized by a membrane-permeant fluorescein derivative with two As(III) substituents, FlAsH, which fluoresces only after the arsenics bind to the cysteine thiols. We now report kinetics and dissociation constants ( approximately 10(-11) M) for FlAsH binding to model tetracysteine peptides. Affinities in vitro and detection limits in living cells are optimized with Xaa-Xaa = Pro-Gly, suggesting that the preferred peptide conformation is a hairpin rather than the previously proposed alpha-helix. Many analogues of FlAsH have been synthesized, including ReAsH, a resorufin derivative excitable at 590 nm and fluorescing in the red. Analogous biarsenicals enable affinity chromatography, fluorescence anisotropy measurements, and electron-microscopic localization of tetracysteine-tagged proteins.
Article
SUMMARYA method to measure the degree of co‐localization of objects in confocal dual‐colour images has been developed. This image analysis produced two coefficients that represent the fraction of co‐localizing objects in each component of a dual‐channel image. The generation of test objects with a Gaussian intensity distribution, at well‐defined positions in both components of dual‐channel images, allowed an accurate investigation of the reliability of the procedure. To do that, the co‐localization coefficients were determined before degrading the image with background, cross‐talk and Poisson noise. These synthesized sources of image deterioration represent sources of deterioration that must be dealt with in practical confocal imaging, namely dark current, non‐specific binding and cross‐reactivity of fluorescent probes, optical cross‐talk and photon noise. The degraded images were restored by filtering and cross‐talk correction. The co‐localization coefficients of the restored images were not significantly different from those of the original undegraded images. Finally, we tested the procedure on images of real biological specimens. The results of these tests correspond with data found in the literature. We conclude that the co‐localization coefficients can provide relevant quantitative information about the positional relation between biological objects or processes.
Article
The process of HIV-1 assembly and release, which occurs largely at the plasma membrane, is driven by the Gag precursor protein, Pr55Gag. Because Gag plays a central role in orchestrating the late stages of the virus replication cycle, elucidating the molecular mechanism by which Gag traffics to the site of virus assembly and defining the cellular machinery that regulates Gag localization are crucial to understanding HIV-1 replication. Ultimately, this information will be vital to devising effective approaches to controlling this pathogen. Numerous studies in recent years have provided new insights into both the viral and cellular determinants of HIV-1 Gag trafficking. The matrix domain of Gag plays an important role in directing Gag to the plasma membrane, in part by directly interacting with a specific phosphoinositide. Several cellular proteins have also been reported to function in Gag trafficking pathways. Current efforts in the field are directed toward understanding how these factors function in promoting Gag transport and virus assembly.