Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Cell Host Microbe. Author manuscript; available in PMC 2013 Oct 18.
Published in final edited form as:
PMCID: PMC3619430
NIHMSID: NIHMS412453
PMID: 23084911

Innate immune sensing of HIV-1 by dendritic cells

Abstract

HIV-1-specific antibodies and CD8+ cytotoxic T cells are detected in most HIV-1-infected people. Yet, HIV-1 infection is not eradicated. Contributing to the failure to mount a sterilizing immune response may be the inability of antigen-presenting dendritic cells (DCs) to sense HIV-1 during acute infection, and thus unable to effectively prime naïve, HIV-1-specific T cells. Recent findings related to DC-expressed innate immune factors, including SAMHD1, TREX1, and TRIM5, provide a molecular basis for understanding why DCs fail to adequately sense invasion by this deadly pathogen and suggest experimental approaches to improve T cell priming to HIV-1 in prophylactic vaccination protocols.

Currently, more than 30 drug formulations are approved that block HIV-1 replication and prevent infected individuals from progressing to AIDS (www.fda.gov). However, treatment with these drugs is life-long and plagued with toxicity, virus drug-resistance, and significant economic cost. According to recent figures from UNAIDS, 34 million people are infected with HIV-1, and for each person who starts anti-HIV-1 drug treatment, it is estimated that there are two to three new infections. In 2010, 1.8 million people died of AIDS-related illnesses and 2.6 million became infected with HIV-1. Efforts to control the AIDS pandemic would benefit from an effective anti-HIV-1 vaccine but, with perhaps one exception (Rerks-Ngarm et al., 2009), attempts to prevent new HIV-1 infection in human vaccine trials have been unsuccessful.

The eradication of smallpox and the effective control of poliovirus, measles, mumps, rubella, and yellow fever offer stark contrast to the public-health experience with HIV-1. The live virus preparations used to immunize against these pathogens were developed empirically, without understanding the mechanisms that underlie the anamnestic response. The success of the vaccines against these viruses - particularly those vaccines that replicate within the host – demonstrates that life-long, protective immune responses can be elicited by vaccination.

In contradistinction, people infected with HIV-1 progress to AIDS despite measurable humoral and cellular immune responses to HIV-1 (Virgin and Walker, 2010). Worse still, HIV-1-infected people with documented, broad anti-HIV-1 immune responses can be secondarily infected with HIV-1 (Altfeld et al., 2002; Smith et al., 2005). Interestingly, failure to protect against reinfection is also seen with Hepatitis C virus (Blackard and Sherman, 2007) and no vaccine is available for this virus either. These observations do not mean that an HIV-1 prophylactic vaccine is impossible, especially given that super-infection with HIV-1 might be aided by the immune dysfunction associated with prior HIV-1 infection. Nonetheless, these observations demonstrate that the immune response targeting HIV-1 differs fundamentally from that against the viruses described above, and suggest that, in the absence of some fundamental modification in vaccine-design, even a live vaccine would be unlikely to alter the outcome of an HIV-1 challenge.

While failure to elicit protective immunity distinguishes HIV-1 (and HCV) from pathogenic viruses such as poliovirus and measles, these are not the only viruses that have eluded efforts to develop a protective vaccine. Dengue infects 500 million people each year, two million of whom suffer complications of hemorrhagic fever (Beatty et al., 2010). Like HIV-1, there are multiple types of Dengue virus, there is no good animal model, and no simple correlates of immunologic protection. The first protective Dengue vaccine - albeit with 30% efficacy - was only possible recently, after 50 years of research (Sabchareon et al., 2012). These results are remarkably similar to the reported 31% efficacy in RV144, the only successful HIV-1 prophylactic vaccine trial (Rerks-Ngarm et al., 2009).

Respiratory Syncytial Virus (RSV) causes >100,000 hospitalizations for pneumonia each year in the United States (Welliver, 2003). As with HIV-1, RSV-specific immune responses are detectable after infection, but protection against recurrent infection is not conferred. Attempts to vaccinate against RSV even caused a paradoxical increase in disease severity, perhaps because the particular adjuvant used – alum - elicited a non-protective CD4+ TH2 response (Lindell et al., 2011).

Ultimately, any advance in mechanistic understanding of protective immunity targeting HIV-1 would likely benefit attempts to control Dengue virus, RSV, and other viruses as well.

This review will assess the large body of literature on HIV-1 and construct a model to explain why the human immune system fails to eliminate or prevent HIV-1 infection. In large part, it will focus on recent developments regarding host cell restriction factors, and attempt to link these findings to what is known about innate immune detection of HIV-1, T cell priming by DCs, and HIV-1 vaccine development.

What permits HIV-1 to escape control by the human immune system?

Several hypotheses offer plausible explanation for HIV-1 persistence in the face of apparent antiviral immunity. Soon after establishment of infection by one or a few HIV-1 virions, variants are selected in response to pressure from HIV-1-specific cytotoxic T lymphocytes (CTL) and neutralizing antibody (Kearney et al., 2009; Henn et al., 2012). It should be noted that high mutagenic rates and complex mixtures of sequence variants known as quasi-species are also well described for other RNA viruses, including some viruses for which there are effective vaccines. The pathogenesis of poliovirus, for example, requires an error-prone viral polymerase and the quasi-species that it generates (Vignuzzi et al., 2006). This suggests that it is not quasi-species per se that distinguishes HIV-1 from other viruses, but rather HIV-1’s capacity to replicate in the face of enormous sequence variation.

The HIV-1 Env glycoprotein, in particular, is distinguished by great sequence diversity and conformational flexibility, as well as an extensive glycan shield and viral entry receptor binding sites that are well concealed and only accessible to antibodies that have undergone extensive somatic mutation (Zhou et al., 2007; Scheid et al., 2009; Bar et al., 2012). Details about this extraordinary molecule and how broadly neutralizing antibodies might be generated by prophylactic vaccination is discussed in an accompanying review by Dennis Burton and colleagues.

Another intuitive explanation for the inability of the immune system to clear infection is that HIV-1 inactivates the same cells of the immune system that protect against viral infection. But tropism for immune cells, as well as the immunosuppression that results, is shared with viruses such as the Measles virus, for which an effective vaccine exists (Tatsuo et al., 2000). Perhaps what is critical for HIV-1 persistence is the exact subclass of immune cell targeted by HIV-1 (Chomont et al., 2009) or the selective targeting of HIV-1-specific CD4+ T cells (Douek et al., 2002).

Importantly, like all retroviruses, HIV-1 integrates its genome into host cell chromosomal DNA, where the virus becomes a permanent genetic element in the infected cell, and in all daughter cells. This enables it to establish a reservoir of HIV-1 infected cells that persist indefinitely, even in the face of an otherwise effective immune response. In contrast, Measles virus cannot integrate, it cannot establish a latent reservoir of infection, and it cannot persist after the Measles virus-specific immune response is established.

Histone deacetylase (HDAC) inhibitors activate latent HIV-1 provirus transcription and have been touted as a means to purge the latent reservoir of HIV-1-infected cells (Archin et al., 2012). The underlying idea is that, upon transcriptional activation of latent HIV-1 proviruses, HIV-1 proteins will be synthesized that would render these infected cells detectable by HIV-1-specific cytotoxic T cells. However, upon activation of transcription by the HDAC inhibitors, the HIV-1-infected cells were not killed by autologous, HIV-1-specific CTL unless the CTL were stimulated with cognate antigen prior to provirus reactivation (Shan et al., 2012). These experiments demonstrate that the HIV-1-specific CTL responses in these individuals were somehow defective.

DC maturation and priming of naïve T cells

Recent research on HIV-1 interaction with dendritic cells (DCs) suggests another hypothesis to explain HIV-1 persistence. DCs are highly heterogeneous, antigen-presenting cells, that initiate acquired immune responses by priming naïve, antigen-specific T cells (Steinman and Idoyaga, 2010). The nature of the T cell response that ensues, whether tolerance or sterilizing antiviral immunity, is determined by the maturation status of the antigen-presenting DC and the array of signals that the DC provides to the naïve, antigen-specific T cell (Figure 1). Among the molecules relevant to T cell priming that DCs elaborate in response to maturation are cell surface MHC Class II, CD80 and CD86, and cytokines such as IL-12 that promote TH1 responses and the cytotoxic T lymphocytes that clear virus-infected cells (Reis e Sousa, 2006; Altfeld et al., 2011).

An external file that holds a picture, illustration, etc.
Object name is nihms412453f1.jpg
Dendritic cells provide multiple signals to naïve T cells

The innate immune system instructs the acquired immune system via a diverse set of signals that dendritic cells provide to naïve, antigen-specific T cells. A range of outcomes are possible - from tolerance to potent antiviral immunity - depending upon the exact signals provided. Protective antiviral immunity requires that naïve T cells specific for viral antigens receive 3 signals from DCs (Reis e Sousa, 2006). Elaboration of the three signals requires DC maturation, most effectively attained by activation of cell-intrinsic pattern recognition receptors. The first signal involves antigen presentation via MHC. The second signal is provided by cell surface molecules that include CD80 and CD86. The third signal is delivered by cytokines that include IL-12, the structure of which is shown (PDB:1F45) (Yoon et al., 2000); the two IL-12 subunits are distinguished from each other by red (IL12A, p35) and blue (IL12B, p40).

DCs mature when pattern recognition receptors (PRRs) are stimulated by pathogen associated molecular patterns (PAMPS) (Akira et al., 2006). Viruses are obligate intracellular parasites, largely dependent upon host cell machinery for replication. They therefore do not generate completely foreign molecules like lipopolysaccharide that distinguish them from the host. Instead, PRRs alert cells to the presence of viruses by detecting more subtle features such as structured replication intermediates, modified nucleic acids, or viral replication complexes in cellular compartments where nucleic acids are not normally found.

Viral nucleic acids can be detected via cell extrinsic or cell intrinsic mechanisms (Iwasaki and Medzhitov, 2010). The former include TLR activation after phagocytosis of virus-infected, apoptotic cells. Such extrinsic PRRs are not expressed by all cell types but are generally restricted to antigen-presenting or phagocytic cells, such as dendritic cells. Cell intrinsic mechanisms detect viral nucleic acid within the infected cell, and the PRRs that detect these PAMPs are generally expressed on a broad range of cell types, including fibroblasts. RIG-I and MDA-5 are cell intrinsic cytosolic receptors that detect structural features unique to viral RNA. A cell intrinsic mechanism also exists for detecting viral DNA in the host cell cytoplasm. Cytosolic sensors for DNA include IFI16, DDX41, DAI, LSm14A, and AIM2 (Unterholzner et al., 2010; Sharma and Fitzgerald, 2011; Sharma et al., 2011; Zhang et al., 2011; Li et al., 2012; Upton et al., 2012).

Optimal priming of naïve T cells to generate potent CD4+TH1 and CD8+ cytotoxic T cell responses requires that PAMP-recognition by PRRs occur within the same DCs that present the antigen, at least under particular experimental conditions (Spörri and Reis e Sousa, 2005; Hou et al., 2008; Kratky et al., 2011). Separation of antigen presentation from PAMP recognition and DC maturation may result in expansion of antigen-specific T cells that lack potent antiviral activity. It is in fact well described that HIV-1-specific T cells express elevated levels of inhibitory molecules such as PD-1, TRAIL, and CTLA-4 (Day et al., 2006; Trautmann et al., 2006; Wherry et al., 2007).

Complicating the assessment of these inhibitory molecules on the HIV-1-specific CD8+ T cells is that these same molecules are up-regulated on activated T cells. Co-expression of PD-1 with another negative regulator of T cell activation, CD160, distinguishes a subset of ineffective HIV-1-specific CD8+ T cells from activated T cells (Peretz et al., 2012). While these ineffective, HIV-1-specific, CD8+ T cells are thought to reflect a state of immune exhaustion that arises during chronic infection with HIV-1 (Khaitan and Unutmaz, 2011), similar phenotypes are observed in short-term experiments ex vivo and in humanized mice (Brainard et al., 2009; Lubong Sabado et al., 2009; Che et al., 2010), suggesting that T cell priming to HIV-1 antigens in the absence of optimal DC maturation might contribute to these ineffective, anti-HIV-1 immune responses.

Innate immune detection of retroviruses

Innate immune detection of viral nucleic acid generally activates type 1 interferon (IFN) and a large number of IFN-stimulated genes with a range of antiviral effector functions. Type 1 IFN also promotes DC maturation and contributes to potent antiviral T cell responses (Longhi et al., 2009). Retroviruses normally do not induce type 1 IFN, perhaps because the level of viral nucleic acid is kept at low level, or because the retrovirus genome is inaccessible to intrinsic PRRs (Fonteneau et al., 2004; Smed-Sörensen et al., 2004; Beignon et al., 2005; Yan et al., 2010; Liberatore and Bieniasz, 2011). This has made it difficult to determine if the innate immune system is capable of detecting retroviruses, and explains why the study of innate immunity against retroviruses has lagged behind that of other viruses. Unlike other viruses that do induce type 1 IFN, retroviruses are generally believed not to encode specific proteins that block innate immune signaling, although some recent reports suggest that HIV-1 disrupts RIG-I and a downstream transcription factor, interferon regulatory factor (IRF3) (Solis et al., 2011; Doehle et al., 2012a; 2012b). Further, under particular experimental conditions in vitro, aborted HIV-1 reverse transcripts in CD4+ T cells have been shown to activate apoptosis and inflammation in a process that involves caspase-3 and caspase-1 (Doitsh et al., 2010).

Important information concerning innate immune detection of retroviruses has been obtained by studying rare instances where adaptive immune control of retroviral infection has been observed. Specific strains of inbred mice such as I/LnJ and C57BL/6J, for example, are resistant to infection with the murine retroviruses Murine Leukemia Virus (MuLV) and Mouse Mammary Tumor Virus (MMTV) via both humoral and cellular mechanisms. This experimental system offers an opportunity to identify the innate immune PRRs that contribute to these protective, anti-retroviral acquired immune responses. The protective antibody responses to both viruses in these mice was TLR7-dependent (Kane et al., 2011b), but induction of protective CTL responses was dependent on a separate, yet to be defined, innate immune detection pathway (Browne and Littman, 2009; Kane et al., 2011b).

The importance of innate immune detection of retroviruses for subsequent acquired immunity was also demonstrated in experiments concerning the influence of the route of infection. When transmitted to pups via ingestion of mother’s milk, MMTV acquires a coat of LPS derived from intestinal bacteria (Kane et al., 2011a). Upon engulfment of these virions by the host cells, TLR4 signaling was activated by the virion-associated LPS. The subsequent cascade of signals resulted in production of IL-10, a tolerogenic cytokine that permitted chronic infection. When MMTV was ingested following sterilization of gut bacteria, or when MMTV infection was initiated parenterally, tolerance to the virus was not observed, and chronic infection did not ensue.

HIV-1 sensing by DCs

Several types of DCs have been described, though they can be divided into two major groups, the myeloid CD11c+ conventional DC (cDC) and the plasmacytoid DC (pDC) (Steinman and Idoyaga, 2010; Altfeld et al., 2011). The pDC, in contrast to the cDC, constitutively expresses TLR7 and TLR9 (Kadowaki et al., 2001). The pDC endocytoses HIV-1 virions via Env interaction with CD4, and the virion genomic RNA is detected by TLR7 within the endocytic compartment (Fonteneau et al., 2004; Beignon et al., 2005; Smed-Sörensen et al., 2005); this innate immune detection of HIV-1 does not require viral replication and, yet, results in the production of copious amounts of type 1 IFN. A clear role for the pDC in priming of naïve, HIV-1-specific CD4+ T cells has not been established, but the IFN secreted by pDC in response to HIV-1 promotes cDC maturation and NK cell activation (Fonteneau et al., 2004; Romagnani et al., 2005). The complex interplay between NK cells, T cells, and DCs, as well as the direct effects of NK cells on HIV-1-infected cells, has been discussed extensively in a recent review (Altfeld et al., 2011).

The importance of the cDC for priming of naïve, anti-HIV-1 T cells is better established than it is for the pDC. HIV-1-pulsed, monocyte-derived DCs, the most common experimental source of human cDCs, prime naïve CD4+ and CD8+ T cells in vitro to generate broad, poly-functional, anti-HIV-1 responses targeting epitopes that mirror those detected after acute infection in vivo (Lubong Sabado et al., 2009). Humanized mice challenged with HIV-1 elaborate similar HIV-1-specific CD4+ and CD8+ T cell responses (Brainard et al., 2009). But, as has been well described in HIV-1 infected people (Day et al., 2006; Trautmann et al., 2006; Wherry et al., 2007), the HIV-1-specific T cells elicited in these assays express elevated levels of inhibitory molecules such as PD-1, TRAIL, and CTLA-4 that suppress the priming of other naïve T cells in a contact-dependent manner (Brainard et al., 2009; Che et al., 2010).

In contrast to pDCs, cDCs challenged with HIV-1 do not make type 1 IFN, and, additionally, they do not produce IL-12 (Fonteneau et al., 2004; Smed-Sörensen et al., 2004; Beignon et al., 2005). The type 1 IFN produced by pDCs in response to challenge with HIV-1 promotes cDC maturation in trans (Fonteneau et al., 2004), but, as mentioned above, if PAMP recognition and antigen-presentation do not occur within the same cell, the critical signals needed for full maturation of cDCs and antiviral immunity may be lacking (Spörri and Reis e Sousa, 2005; Hou et al., 2008; Kratky et al., 2011). These observations demonstrate that cell intrinsic, innate immune sensing of HIV-1 is insufficient to effect complete cDC maturation. Given the clear importance of the cDC for T cell priming, the discussion that follows below will focus on this DC subtype.

Non-replicative interactions dominate HIV-1 entry into DCs

cDCs fail to mature upon HIV-1 exposure probably because the virus cannot infect these cells (Cameron et al., 1992; Smed-Sörensen et al., 2005). Productive infection by HIV-1 requires virion membrane fusion with the target cell membrane, driven by the interaction between HIV-1 gp120/gp41 and the cell surface entry receptors CD4 and CCR5. This results in delivery of the virion core into the cytoplasm where reverse transcription takes place (McDonald et al., 2002). Levels of cell surface CD4 and CCR5 are relatively low on most DC subsets (Caux et al., 1992; Granelli-Piperno et al., 2006), and when HIV-1 encounters DCs it is more likely to be endocytosed via interaction with DC-SIGN or other C-type lectins (Piguet and Steinman, 2007) (Figure 2). Reverse transcription does not occur within this endocytic compartment, PRRs are not activated, and DCs fail to mature (Granelli-Piperno et al., 2004).

An external file that holds a picture, illustration, etc.
Object name is nihms412453f2.jpg
HIV-1 virion interaction with dendritic cells is dominated by non-productive entry pathways

Productive infection requires that HIV-1 virions gain access to the target cell cytoplasm via HIV-1 glycoprotein (gp120/gp41) interacting with host cell receptors (CD4 and either CCR5 or CXCR4). Cell surface levels of CD4 are low on many types of dendritic cells, and HIV-1 virions are endocytosed via interaction with cell surface lectins such as DC-SIGN. Replication of the HIV-1 genomic RNA (green lines) does not occur within the endocytic compartment and innate immune signaling is not activated within conventional, antigen-presenting DCs. However, the endocytosed virion is degraded and HIV-1 proteins are therefore efficiently presented in the absence of DC maturation. Alternatively, the glycosphingolipid GM3 in the HIV-1 virion membrane is recognized by an unknown receptor on DCs (designated X). This results in virion capture and transfer to susceptible CD4+ T cells via infectious synapses.

HIV-1 virions that are endocytosed by DCs are probably transferred to a lysosomal compartment – and a fraction to proteasomes – and rapidly degraded for processing and presentation in the absence of DC maturation (Moris et al., 2004; 2006). Some virions escape degradation by trafficking to immunologic synapses, from which they are efficiently transferred to highly permissive CD4+ T cells (Pope et al., 1994; McDonald et al., 2003). Avoidance of the degradation pathway and transfer of infectious virions to immunologic synapses is an Env-independent process that requires incorporation of the host cell-derived, glycosphingolipid GM3 into the HIV-1 virion membrane (Izquierdo-Useros et al., 2012; Puryear et al., 2012). The outcome of interaction with the DC C-type lectin and GM3 pathways would be tolerance to HIV-1 and potentiation of HIV-1 spread to highly permissive, activated CD4+ T cells (Figure 2).

Experimentally, HIV-1 virions can be pseudotyped with heterologous viral glycoproteins (Figure 3). This technique can be exploited to redirect HIV-1 transduction to cells that are normally not targeted by HIV-1. Pseudotyping of HIV-1 with the vesicular stomatitis virus glycoprotein (VSV G) drives fusion of HIV-1 virions from endosomes into the cytoplasm and greatly increases the efficiency with which HIV-1 enters the dendritic cell cytoplasm (Granelli-Piperno et al., 2000). This has become the standard methodology for transducing dendritic cells with HIV-1-based vectors (Berger et al., 2011).

An external file that holds a picture, illustration, etc.
Object name is nihms412453f3.jpg
Bypass of the block to productive entry reveals a block to HIV-1 reverse transcription in DCs due to SAMHD1

The block to productive entry can be overcome experimentally by producing HIV-1 virions bearing the vesicular stomatitis virus glycoprotein (VSV G). Once in the dendritic cell cytoplasm, HIV-1 virions fail to reverse transcribe the viral genomic RNA template (green), into viral cDNA (red), because the dNTP concentration is too low to support cDNA synthesis. This block to reverse transcription is due to the dNTP triphosphohydrolase SAMHD1.

SAMHD1 prevents HIV-1 reverse transcription in DCs

Although VSV G pseudotyping increases the efficiency with which HIV-1 virions enter DCs, a second block to replication occurs at the level of reverse transcription (Figure 3). The first clue to the nature of this block came from the observation that viruses of the human immunodeficiency virus type 2 (HIV-2) and the simian immunodeficiency SIVsm/SIVmac lineage are more efficient than HIV-1 at transducing DCs, and that transduction by SIVMAC vectors required the vpx accessory gene (Mangeot et al., 2002). When provided in trans, SIV virus-like particles (VLPs) greatly increased the efficiency of HIV-1 reverse transcription and transduction in DCs (Goujon et al., 2006). This increase in HIV-1 infectivity also required the SIV Vpx protein (Goujon et al., 2006) and was mimicked by proteasome inhibitors (Goujon et al., 2007). Vpx was then shown to associate with the CUL4A E3 ubiquitin ligase complex via direct binding to DCAF1 (Sharova et al., 2008; 2008; Srivastava et al., 2008; 2008; Bergamaschi et al., 2009; 2009) and heterokaryon experiments demonstrated the presence of a dominant-acting inhibitor in myeloid cells that was sensitive to Vpx (Sharova et al., 2008). The observation that myeloid cells have 100 times lower dNTP pools than do HIV-1-permissive, activated CD4+ T cells (Diamond et al., 2004; Kennedy et al., 2010) was an additional clue to the mechanism of the block to reverse transcription since reverse transcription requires dNTPs for synthesis of viral cDNA.

Vpx was then shown to promote the degradation of SAMHD1 (Goldstone et al., 2011; Powell et al., 2011; Lahouassa et al., 2012), a triphosphohydrolase that converts deoxynucleoside triphosphates (dNTPs) to the constituent deoxynucleosides and inorganic triphosphate (Goldstone et al., 2011; Powell et al., 2011; Lahouassa et al., 2012). It was reasonable to hypothesize that, by depleting dNTPs, SAMHD1 would block reverse transcription. Indeed, the kinetics of SAMHD1 degradation that followed Vpx delivery correlated with rising dNTP levels, and subsequent increase in HIV-1 reverse transcription (Kim et al., 2012). SAMHD1 knockdown permitted HIV-1 reverse transcription to proceed in myeloid cells in the absence of Vpx (Hrecka et al., 2011; Laguette et al., 2011). Cells from Aicardi-Goutières Syndrome (AGS) patients that bear mutations in SAMHD1 were similarly permissive for HIV-1 in the absence of Vpx. This inflammatory disease will be discussed further below. The antiviral effect of SAMHD1 has also been reported more recently in non-dividing, resting CD4+ T cells (Baldauf et al., 2012).

Consequences of DC transduction

Efficient HIV-1 transduction of DCs can be achieved by pseudotyping particles with VSV G in the presence of Vpx-bearing VLPs or exogenous nucleosides. One group has reported that DC maturation can occur in response to such high-efficiency transduction by HIV-1, and that DC maturation was associated with increased responsiveness of HIV-1-specific T cell clones (Manel et al., 2010). DC maturation required interaction between the cellular protein cyclophilin A and HIV-1 capsid protein synthesized de novo from the provirus (Manel et al., 2010). Cyclophilin A is a well characterized HIV-1 capsid-binding protein that regulates HIV-1 infectivity (Sokolskaja and Luban, 2006).

Other groups have failed to detect maturation of DCs after HIV-1 transduction (Fonteneau et al., 2004; Granelli-Piperno et al., 2004; Beignon et al., 2005; Smed-Sörensen et al., 2005; Granelli-Piperno et al., 2006; Pertel et al., 2011a; 2011b), even when transduction was pushed to high levels with Vpx. Perhaps this discrepancy is due to variation in additional host factors that regulate HIV-1 replication in DCs, such as TREX1 (discussed below), or other myeloid-specific, innate immune factors that are yet to be discovered. For example, Vpx rescues HIV-1 transduction of DCs from the antiviral state established by prior treatment with exogenous type 1 IFN via a mechanism that is independent of SAMHD1 and DCAF1-CUL4A. (Pertel et al., 2011b).

Infection with HIV-2, a virus that expresses Vpx, is less likely to cause AIDS than HIV-1 (de Silva et al., 2008). Perhaps by permitting HIV-2 to complete reverse transcription within DCs, Vpx enables cell-intrinsic PRRs within these cells to sense HIV-2. Signaling from the PRRs would direct the antigen-presenting DCs to mature and thus render acquired immune responses against HIV-2 more effective. Consistent with this idea, polyfunctional, virus-specific, T cell responses are more commonly observed with HIV-2 than with HIV-1 (Duvall et al., 2008). Further evidence in support of this model comes from experiments in which an HIV-1 provirus was modified to express and package SIV Vpx; the resulting virus infected myeloid cells more efficiently than did the parental HIV-1 virus (Sunseri et al., 2011). Additionally, type 1 IFN was detected in these spreading infections (Sunseri et al., 2011), something that was not seen with wild-type HIV-1.

The correlation between Vpx and immune control is not perfect, in that SIVMAC239, another virus that encodes Vpx, is highly pathogenic in a non-native host monkey species. However, SIVSM the naturally occurring virus from which SIVMAC239 was derived, is not pathogenic in its native host (Chahroudi et al., 2012).

TREX1 conceals HIV-1 cDNA from detection by the innate immune system

The presence of DNA in the cytoplasm of DCs and other cell types, whether from a transfected plasmid, synthetic dsDNA, intracellular bacteria, or infection with a DNA virus, activates type 1 IFN (Sharma and Fitzgerald, 2011). The nuclease TREX1 was identified in a biochemical screen intended to determine the identity of the PRR that activates type 1 IFN in response to DNA within the cytoplasm of macrophages (Stetson et al., 2008). However, in the follow-up functional test, TREX1 disruption did not block type 1 IFN induction by DNA, demonstrating that TREX1 was not the sought-after PRR. Paradoxically, TREX1-deficient animals had elevated type 1 IFN and presented with inflammatory myocarditis (Stetson et al., 2008). Additionally, cDNA from endogenous retroelements was increased in TREX1-deficient cells, consistent with a role for TREX1 in suppressing retroviral cDNA accumulation and blocking cell intrinsic detection of DNA (Stetson et al., 2008). Amazingly, reverse transcriptase inhibitors that prevent synthesis of retroviral cDNA limited the inflammatory disease in TREX1-null mice (Beck-Engeser et al., 2011).

Subsequently it was reported that when TREX1 is disrupted, challenge with HIV-1 resulted in elevated levels of HIV-1 cDNA and type 1 IFN induction (Yan et al., 2010). This study demonstrated that HIV-1 cDNA can be sensed by the innate immune system via an intrinsic mechanism. However, the cDNA resulting from HIV-1 reverse transcription does not normally accumulate to sufficient levels to be detected unless TREX1 is disrupted (Figure 4). The PRR responsible for detecting HIV-1 cDNA under these conditions has not been determined, although the cytosolic viral DNA sensor IFI16 is a possible candidate (Unterholzner et al., 2010; Sharma et al., 2011).

An external file that holds a picture, illustration, etc.
Object name is nihms412453f4.jpg
The human exonuclease TREX1 limits the level of HIV-1 cDNA

HIV-1 reverse transcriptase generates a cDNA (red lines) from the HIV-1 genomic RNA (green) in the target cell cytoplasm. Normally, the HIV-1 cDNA is not detected by pattern recognition receptors (PRRs) that recognize DNA in the cytoplasm. When the host cell TREX1 nuclease is disrupted, either by mutation or by siRNA, HIV-1 cDNA accumulates to supra-normal levels and activates type 1 IFN. Though several candidate receptors exist (e.g., IFI16), the cell intrinsic PRR that recognizes HIV-1 cDNA under these conditions has not been identified.

Mutations in the gene encoding TREX1, as well as in SAMHD1, the HIV-1 restriction factor specific to dendritic cells, macrophages, and nondividing T cells, are associated with Aicardi-Goutières Syndrome (AGS). This rare, lethal, inflammatory syndrome is characterized by elevated type 1 IFN (Crow and Rehwinkel, 2009; Rice et al., 2009). AGS-associated mutations are also found in genes encoding each of the three proteins that constitute the enzyme RNASEH2 (Crow et al., 2006). This enzyme cleaves ribonucleotides within RNA-DNA duplexes like those found in the intermediate products of reverse transcription. Whether RNASEH2 influences innate immune recognition of HIV-1 in DCs remains to be determined. The current model for AGS pathophysiology, whether due to mutations in TREX1, SAMHD1, or RNASEH2, is that inflammation results from activation of cytosolic PRRs in response to the increased levels of endogenous reverse transcripts or reaction intermediates that accumulate when either of these factors is mutated. By demonstrating that the human innate immune system is indeed capable of detecting HIV-1, the AGS factors offer a new paradigm in AIDS vaccine research.

TRIM5, CA-specific restriction, and innate immune signaling in DCs

TRIM5 was identified in expression screens seeking the factor responsible for an anti-HIV-1 activity in diverse cell types from macaques and owl monkeys (Sayah et al., 2004; Stremlau et al., 2004). The TRIM5 orthologues from these two non-human primates were then shown to potently block HIV-1 transduction (10–100-fold). The viral determinant for sensitivity to TRIM5 restriction is the capsid protein (CA), but TRIM5 binds to the capsid lattice, not free capsid protein (Grütter and Luban, 2012) (Figure 5) and binding induces TRIM5 to form a complementary lattice (Ganser-Pornillos et al., 2011). Because the TRIM5-CA interaction is so complicated, a robust binding assay has not yet been developed, and conclusions about binding strength are extrapolated from restriction activity assays or non-quantitative, particulate association assays (Stremlau et al., 2006).

An external file that holds a picture, illustration, etc.
Object name is nihms412453f5.jpg
TRIM5 is a pattern recognition receptor specific for the retrovirus capsid lattice

The cytoplasmic E3 ubiquitin ligase TRIM5 exists as a dimer in the cytoplasm where it signals weakly or not at all. Upon infection with a retrovirus, if avidity for the capsid lattice is great enough, TRIM5 multimerizes to form a complementary lattice (Ganser-Pornillos et al., 2011) and E3 ubiquitin ligase activity is increased (Pertel et al., 2011a). In complex with the heterodimeric E2 (UBC13/UEV1A), TRIM5 synthesizes unattached K63-linked ubiquitin chains that activate the TAK1 kinase complex and innate immune signaling (Pertel et al., 2011a). Most HIV-1 strains (blue capsid) are only weakly detected by the human TRIM5 orthologue. HIV-1 capsid variants that are better recognized by human TRIM5 (orange capsid) would increase innate immune signaling, and might make better immunogens.

Phylogenetic comparisons have shown that the sequences encoding the capsid binding domain of TRIM5 are evolving at rates faster than any coding sequence in the primate genome (Sawyer et al., 2005). It has been suggested that this positive selection is driven by challenges in the remote past from pathogenic retroviruses. The human TRIM5 orthologue weakly inhibits lab strains of HIV-1, on the order of two-fold, in quantitative, single-cycle infectivity assays (Sokolskaja et al., 2006; Battivelli et al., 2010; 2011). The weak inhibition of HIV-1 by human TRIM5 is remedied by replacing a few capsid binding residues in the human orthologue with those from the macaque (Stremlau et al., 2005). It is possible that human TRIM5 is an active restriction factor in vivo and that it just does not recognize lab strains of HIV-1 very well. One group has recently cloned HIV-1 sequences directly from infected people and found impressive variation in TRIM5 sensitivity. Some HIV-1 isolates were 10 times more sensitive than the lab strain standard (Battivelli et al., 2010; 2011). Interestingly, the variant amino acids in these TRIM5-sensitive strains were located in dominant CTL epitopes within capsid, suggesting that escape from CTL forced the acquisition of increased TRIM5 sensitivity.

TRIM5 promotes innate immune signaling (Tareen and Emerman, 2011; Pertel et al., 2011a) via a mechanism that involves synthesis of unattached K63-linked ubiquitin chains and activation of the TAK1 kinase complex (Pertel et al., 2011a) (Figure 5). This biochemical activity, and correspondingly the innate immune signaling within DCs, was amplified greatly by infection with retroviruses that bear restriction-sensitive capsids (Pertel et al., 2011a). The results demonstrating that TRIM5 acts as a PRR specific for the retrovirus capsid lattice were obtained using select combinations of TRIM5 orthologues with specific retrovirus capsids (Pertel et al., 2011a). For example, human TRIM5 in DCs or macrophages signaled in response to the restricted N-tropic MLV, but not to the isogenic, non-restricted, B-MLV. Owl monkey TRIM5 signaled in response to restricted HIV-1, but not to the unrestricted SIV.

Induction of cytokines was not detected in human DCs challenged with HIV-1, most probably because the avidity of human TRIM5 for capsid from standard strains of HIV-1 is too weak to activate signaling. For signaling to occur in human dendritic cells in response to HIV-1 capsid, the avidity of the interaction will need to be increased. This might be achieved by using HIV-1 capsid variants with CTL escape mutations that confer greatly increased HIV-1 sensitivity (Battivelli et al., 2010; 2011) (Figure 5). Alternatively, if robust binding assays are developed for measuring the interaction of TRIM5 with soluble capsid lattice components (Zhao et al., 2011), screens might be undertaken to identify small molecules that increase the avidity of the interaction.

Conclusions and perspectives

HIV-1 vaccine development continues to present enormous intellectual and technical challenges, demonstrating the ongoing need for fundamental biological research concerning pathogen immunity. Several well-characterized properties of HIV-1 explain why vaccine development is so difficult. Additionally, many observations suggest that HIV-1 detection by PRRs within DCs is insufficient to mature antigen-presenting DCs, with the result that naïve, HIV-1-specific T cells are not primed for optimal anti-HIV-1 immunity. Progress towards a vaccine will likely require that we understand how conditions can be modified such that HIV-1 is better detected by the innate immune system. Heightened detection of HIV-1 by DCs would result in qualitative improvement in T cell priming, increased survival and effector function of anti-HIV-1 CD8+ T cells, and the requisite affinity maturation for generation of broadly neutralizing anti-HIV-1 antibodies.

Optimal T cell priming likely requires HIV-1 replication within DCs and the resulting activation of intrinsic PRRs. A single block to HIV-1 replication in DCs would be sufficient to preclude a potent acquired immune response. The large number of recently identified blocks to productive HIV-1 infection within these cells, as discussed above, is therefore quite impressive. Expression of DC-SIGN and low levels of CD4 on the surface of DCs, along with GM3-glycosphingolipid within the HIV-1 virion membrane, divert HIV-1 away from productive entry pathways. Reverse transcription is prevented by the nucleotidase SAMHD1. If HIV-1 cDNA is synthesized, the host nuclease TREX1 prevents the cDNA from accumulating to levels that are detectable by cytoplasmic PRRs. Innate immune signaling pathways are not activated by human TRIM5 unless CTL forces the HIV-1 capsid to mutate such that the avidity of TRIM5 for the capsid protein lattice is increased.

This fundamental biological information, along with technical advances in molecular biology, offer experimental approaches to improve DC maturation in response to HIV-1 infection. Non-productive entry pathways can be overcome by pseudotyping HIV-1 virions with VSV G. dNTP levels necessary for reverse transcription can be achieved by knockdown of SAMHD1, delivery of Vpx, or deoxynucleosides. TREX1 knockdown permits HIV-1 cDNA to accumulate to sufficient levels to activate cytoplasmic PRRs. HIV-1 capsid variants that better engage TRIM5 can potentiate the innate immune signaling elicited by the other measures.

Going forward, it will be necessary to determine which combination of the interventions described above allows maximal DC maturation and requisite priming of naïve anti-HIV-1 T cells. These experiments could be completed in tissue culture using human blood cells or using humanized mouse models, and may provide tools and insights that could be translated to immunize humans. For example, autologous DCs treated ex vivo with HIV-1 vectors, immunogens, or drugs might be re-injected as a vaccine. Simpler vaccine formulations would of course be preferable and will hopefully spring from the information these experiments provide.

Going forward, valuable information that might come from these experiments would include the identification of the PRR recognizing HIV-1 cDNA; this PRR would be an appealing cellular target for new adjuvant development. Other than IL-12, what are the specific factors produced by mature, antigen-presenting DCs that optimally prime naïve T cells for protection against HIV-1? What is the identity of the glycosphingolipid GM3 receptor in DCs and would its blockade increase the immunogenicity of HIV-1? Would knockdown of RNASEH2, a gene that is mutated in Aicardi-Goutieres syndrome, increase the level of HIV-1 reverse transcription intermediates and activate type 1 IFN, as is seen with TREX1 and HIV-1 cDNA? What is the minimal structure of the retrovirus capsid lattice that is recognized by TRIM5, and could this information be exploited to improve HIV-1 CA detection by the human TRIM5 orthologue? Ultimately, it will be of great interest to know if new vaccines that increase the intrinsic detection of HIV-1 within antigen-presenting DCs reduce the frequency of new HIV-1 infection in populations at risk for HIV-1.

Acknowledgments

This review was written in fond memory of a very inspiring teacher, Ralph Steinman. Thanks to Tatyana Golovkina for critical reading of the manuscript and to Rahm Gummuluru, Rick Koup, Harmit Malik, Douglas Nixon, and Celia Schiffer for informative discussions. Apologies to the many investigators whose discoveries and influence is not adequately acknowledged due to space limitations or my ignorance. The author is funded by grants from the Swiss National Science Foundation, NIAID (USA), and a NIDA Avant-Garde Award.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • Akira S, Uematsu S, Takeuchi O. Pathogen recognition and innate immunity. Cell. 2006;124:783–801. [PubMed] [Google Scholar]
  • Altfeld M, Allen TM, Yu XG, Johnston MN, Agrawal D, Korber BT, Montefiori DC, O’Connor DH, Davis BT, Lee PK, et al. HIV-1 superinfection despite broad CD8+ T-cell responses containing replication of the primary virus. Nature. 2002;420:434–439. [PubMed] [Google Scholar]
  • Altfeld M, Fadda L, Frleta D, Bhardwaj N. DCs and NK cells: critical effectors in the immune response to HIV-1. Nat Rev Immunol. 2011;11:176–186. [PMC free article] [PubMed] [Google Scholar]
  • Archin NM, Liberty AL, Kashuba AD, Choudhary SK, Kuruc JD, Crooks AM, Parker DC, Anderson EM, Kearney MF, Strain MC, et al. Administration of vorinostat disrupts HIV-1 latency in patients on antiretroviral therapy. Nature. 2012;487:482–485. [PMC free article] [PubMed] [Google Scholar]
  • Baldauf H-M, Pan X, Erikson E, Schmidt S, Daddacha W, Burggraf M, Schenkova K, Ambiel I, Wabnitz G, Gramberg T, et al. SAMHD1 restricts HIV-1 infection in resting CD4(+) T cells. Nat Med 2012 [PMC free article] [PubMed] [Google Scholar]
  • Bar KJ, Tsao CY, Iyer SS, Decker JM, Yang Y, Bonsignori M, Chen X, Hwang KK, Montefiori DC, Liao HX, et al. Early low-titer neutralizing antibodies impede HIV-1 replication and select for virus escape. PLoS Pathog. 2012;8:e1002721. [PMC free article] [PubMed] [Google Scholar]
  • Battivelli E, Lecossier D, Matsuoka S, Migraine J, Clavel F, Hance AJ. Strain-specific differences in the impact of human TRIM5alpha, different TRIM5alpha alleles, and the inhibition of capsid-cyclophilin A interactions on the infectivity of HIV-1. J Virol. 2010;84:11010–11019. [PMC free article] [PubMed] [Google Scholar]
  • Battivelli E, Migraine J, Lecossier D, Yeni P, Clavel F, Hance AJ. Gag cytotoxic T lymphocyte escape mutations can increase sensitivity of HIV-1 to human TRIM5alpha, linking intrinsic and acquired immunity. J Virol. 2011;85:11846–11854. [PMC free article] [PubMed] [Google Scholar]
  • Beatty ME, Stone A, Fitzsimons DW, Hanna JN, Lam SK, Vong S, Guzman MG, Mendez-Galvan JF, Halstead SB, Letson GW, et al. Best practices in dengue surveillance: a report from the Asia-Pacific and Americas Dengue Prevention Boards. PLoS Negl Trop Dis. 2010;4:e890. [PMC free article] [PubMed] [Google Scholar]
  • Beck-Engeser GB, Eilat D, Wabl M. An autoimmune disease prevented by anti-retroviral drugs. Retrovirology. 2011;8:91. [PMC free article] [PubMed] [Google Scholar]
  • Beignon AS, McKenna K, Skoberne M, Manches O, DaSilva I, Kavanagh DG, Larsson M, Gorelick RJ, Lifson JD, Bhardwaj N. Endocytosis of HIV-1 activates plasmacytoid dendritic cells via Toll-like receptor-viral RNA interactions. J Clin Invest. 2005;115:3265–3275. [PMC free article] [PubMed] [Google Scholar]
  • Bergamaschi A, Ayinde D, David A, Le Rouzic E, Morel M, Collin G, Descamps D, Damond F, Brun-Vezinet F, Nisole S, et al. The human immunodeficiency virus type 2 Vpx protein usurps the CUL4A-DDB1 DCAF1 ubiquitin ligase to overcome a postentry block in macrophage infection. J Virol. 2009;83:4854–4860. [PMC free article] [PubMed] [Google Scholar]
  • Berger G, Durand S, Goujon C, Nguyen XN, Cordeil S, Darlix JL, Cimarelli A. A simple, versatile and efficient method to genetically modify human monocyte-derived dendritic cells with HIV-1-derived lentiviral vectors. Nat Protoc. 2011;6:806–816. [PubMed] [Google Scholar]
  • Blackard JT, Sherman KE. Hepatitis C virus coinfection and superinfection. J Infect Dis. 2007;195:519–524. [PubMed] [Google Scholar]
  • Brainard DM, Seung E, Frahm N, Cariappa A, Bailey CC, Hart WK, Shin HS, Brooks SF, Knight HL, Eichbaum Q, et al. Induction of robust cellular and humoral virus-specific adaptive immune responses in human immunodeficiency virus-infected humanized BLT mice. J Virol. 2009;83:7305–7321. [PMC free article] [PubMed] [Google Scholar]
  • Browne EP, Littman DR. Myd88 is required for an antibody response to retroviral infection. PLoS Pathog. 2009;5:e1000298. [PMC free article] [PubMed] [Google Scholar]
  • Cameron PU, Freudenthal PS, Barker JM, Gezelter S, Inaba K, Steinman RM. Dendritic cells exposed to human immunodeficiency virus type-1 transmit a vigorous cytopathic infection to CD4+ T cells. Science. 1992;257:383–387. [PubMed] [Google Scholar]
  • Caux C, Dezutter-Dambuyant C, Schmitt D, Banchereau J. GM-CSF and TNF-alpha cooperate in the generation of dendritic Langerhans cells. Nature. 1992;360:258–261. [PubMed] [Google Scholar]
  • Chahroudi A, Bosinger SE, Vanderford TH, Paiardini M, Silvestri G. Natural SIV hosts: showing AIDS the door. Science. 2012;335:1188–1193. [PMC free article] [PubMed] [Google Scholar]
  • Che KF, Sabado RL, Shankar EM, Tjomsland V, Messmer D, Bhardwaj N, Lifson JD, Larsson M. HIV-1 impairs in vitro priming of naïve T cells and gives rise to contact-dependent suppressor T cells. Eur J Immunol. 2010;40:2248–2258. [PMC free article] [PubMed] [Google Scholar]
  • Chomont N, Far, El M, Ancuta P, Trautmann L, Procopio FA, Yassine-Diab B, Boucher G, Boulassel M-R, Ghattas G, Brenchley JM, et al. HIV reservoir size and persistence are driven by T cell survival and homeostatic proliferation. Nat Med. 2009;15:893–900. [PMC free article] [PubMed] [Google Scholar]
  • Crow YJ, Rehwinkel J. Aicardi-Goutieres syndrome and related phenotypes: linking nucleic acid metabolism with autoimmunity. Human Molecular Genetics. 2009;18:R130–R136. [PMC free article] [PubMed] [Google Scholar]
  • Crow YJ, Leitch A, Hayward BE, Garner A, Parmar R, Griffith E, Ali M, Semple C, Aicardi J, Babul-Hirji R, et al. Mutations in genes encoding ribonuclease H2 subunits cause Aicardi-Goutières syndrome and mimic congenital viral brain infection. Nat Genet. 2006;38:910–916. [PubMed] [Google Scholar]
  • Day CL, Kaufmann DE, Kiepiela P, Brown JA, Moodley ES, Reddy S, Mackey EW, Miller JD, Leslie AJ, DePierres C, et al. PD-1 expression on HIV-specific T cells is associated with T-cell exhaustion and disease progression. Nature. 2006;443:350–354. [PubMed] [Google Scholar]
  • de Silva TI, Cotten M, Rowland-Jones SL. HIV-2: the forgotten AIDS virus. Trends Microbiol. 2008;16:588–595. [PubMed] [Google Scholar]
  • Diamond TL, Roshal M, Jamburuthugoda VK, Reynolds HM, Merriam AR, Lee KY, Balakrishnan M, Bambara RA, Planelles V, Dewhurst S, et al. Macrophage tropism of HIV-1 depends on efficient cellular dNTP utilization by reverse transcriptase. J Biol Chem. 2004;279:51545–51553. [PMC free article] [PubMed] [Google Scholar]
  • Doehle BP, Chang K, Fleming L, McNevin J, Hladik F, McElrath MJ, Gale M. Vpu deficient HIV strains stimulate innate immune signaling responses in target cells. J Virol 2012a [PMC free article] [PubMed] [Google Scholar]
  • Doehle BP, Chang K, Rustagi A, McNevin J, McElrath MJ, Gale M. Vpu Mediates Depletion of Interferon Regulatory Factor 3 during HIV Infection by a Lysosome-Dependent Mechanism. J Virol. 2012b;86:8367–8374. [PMC free article] [PubMed] [Google Scholar]
  • Doitsh G, Cavrois M, Lassen KG, Zepeda O, Yang Z, Santiago ML, Hebbeler AM, Greene WC. Abortive HIV infection mediates CD4 T cell depletion and inflammation in human lymphoid tissue. Cell. 2010;143:789–801. [PMC free article] [PubMed] [Google Scholar]
  • Douek DC, Brenchley JM, Betts MR, Ambrozak DR, Hill BJ, Okamoto Y, Casazza JP, Kuruppu J, Kunstman K, Wolinsky S, et al. HIV preferentially infects HIV-specific CD4+ T cells. Nature. 2002;417:95–98. [PubMed] [Google Scholar]
  • Duvall MG, Precopio ML, Ambrozak DA, Jaye A, McMichael AJ, Whittle HC, Roederer M, Rowland-Jones SL, Koup RA. Polyfunctional T cell responses are a hallmark of HIV-2 infection. Eur J Immunol. 2008;38:350–363. [PMC free article] [PubMed] [Google Scholar]
  • Fonteneau JF, Larsson M, Beignon AS, McKenna K, DaSilva I, Amara A, Liu YJ, Lifson JD, Littman DR, Bhardwaj N. Human immunodeficiency virus type 1 activates plasmacytoid dendritic cells and concomitantly induces the bystander maturation of myeloid dendritic cells. J Virol. 2004;78:5223–5232. [PMC free article] [PubMed] [Google Scholar]
  • Ganser-Pornillos BK, Chandrasekaran V, Pornillos O, Sodroski JG, Sundquist WI, Yeager M. Hexagonal assembly of a restricting TRIM5alpha protein. Proceedings of the National Academy of Sciences. 2011;108:534–539. [PMC free article] [PubMed] [Google Scholar]
  • Goldstone DC, Ennis-Adeniran V, Hedden JJ, Groom HCT, Rice GI, Christodoulou E, Walker PA, Kelly G, Haire LF, Yap MW, et al. HIV-1 restriction factor SAMHD1 is a deoxynucleoside triphosphate triphosphohydrolase. Nature. 2011;480:379–382. [PubMed] [Google Scholar]
  • Goujon C, Jarrosson-Wuillème L, Bernaud J, Rigal D, Darlix JL, Cimarelli A. With a little help from a friend: increasing HIV transduction of monocyte-derived dendritic cells with virion-like particles of SIV(MAC) Gene Ther. 2006;13:991–994. [PubMed] [Google Scholar]
  • Goujon C, Rivière L, Jarrosson-Wuilleme L, Bernaud J, Rigal D, Darlix JL, Cimarelli A. SIVSM/HIV-2 Vpx proteins promote retroviral escape from a proteasome-dependent restriction pathway present in human dendritic cells. Retrovirology. 2007;4:2. [PMC free article] [PubMed] [Google Scholar]
  • Granelli-Piperno A, Golebiowska A, Trumpfheller C, Siegal FP, Steinman RM. HIV-1-infected monocyte-derived dendritic cells do not undergo maturation but can elicit IL-10 production and T cell regulation. Proc Natl Acad Sci USA. 2004;101:7669–7674. [PMC free article] [PubMed] [Google Scholar]
  • Granelli-Piperno A, Shimeliovich I, Pack M, Trumpfheller C, Steinman RM. HIV-1 selectively infects a subset of nonmaturing BDCA1-positive dendritic cells in human blood. J Immunol. 2006;176:991–998. [PubMed] [Google Scholar]
  • Granelli-Piperno A, Zhong L, Haslett P, Jacobson J, Steinman RM. Dendritic cells, infected with vesicular stomatitis virus-pseudotyped HIV-1, present viral antigens to CD4+ and CD8+ T cells from HIV-1-infected individuals. J Immunol. 2000;165:6620–6626. [PubMed] [Google Scholar]
  • Grütter MG, Luban J. TRIM5 structure, HIV-1 capsid recognition, and innate immune signaling. Curr Opin Virol. 2012;2:142–150. [PMC free article] [PubMed] [Google Scholar]
  • Henn MR, Boutwell CL, Charlebois P, Lennon NJ, Power KA, Macalalad AR, Berlin AM, Malboeuf CM, Ryan EM, Gnerre S, et al. Whole genome deep sequencing of HIV-1 reveals the impact of early minor variants upon immune recognition during acute infection. PLoS Pathog. 2012;8:e1002529. [PMC free article] [PubMed] [Google Scholar]
  • Hou B, Reizis B, DeFranco AL. Toll-like receptors activate innate and adaptive immunity by using dendritic cell-intrinsic and -extrinsic mechanisms. Immunity. 2008;29:272–282. [PMC free article] [PubMed] [Google Scholar]
  • Hrecka K, Hao C, Gierszewska M, Swanson SK, Kesik-Brodacka M, Srivastava S, Florens L, Washburn MP, Skowronski J. Vpx relieves inhibition of HIV-1 infection of macrophages mediated by the SAMHD1 protein. Nature. 2011;474:658–661. [PMC free article] [PubMed] [Google Scholar]
  • Iwasaki A, Medzhitov R. Regulation of adaptive immunity by the innate immune system. Science. 2010;327:291–295. [PMC free article] [PubMed] [Google Scholar]
  • Izquierdo-Useros N, Lorizate M, Contreras FX, Rodriguez-Plata MT, Glass B, Erkizia I, Prado JG, Casas J, Fabriàs G, Kräusslich HG, et al. Sialyllactose in viral membrane gangliosides is a novel molecular recognition pattern for mature dendritic cell capture of HIV-1. PLoS Biol. 2012;10:e1001315. [PMC free article] [PubMed] [Google Scholar]
  • Kadowaki N, Ho S, Antonenko S, Malefyt RW, Kastelein RA, Bazan F, Liu YJ. Subsets of human dendritic cell precursors express different toll-like receptors and respond to different microbial antigens. J Exp Med. 2001;194:863–869. [PMC free article] [PubMed] [Google Scholar]
  • Kane M, Case LK, Kopaskie K, Kozlova A, MacDearmid C, Chervonsky AV, Golovkina TV. Successful transmission of a retrovirus depends on the commensal microbiota. Science. 2011a;334:245–249. [PMC free article] [PubMed] [Google Scholar]
  • Kane M, Case LK, Wang C, Yurkovetskiy L, Dikiy S, Golovkina TV. Innate Immune Sensing of Retroviral Infection via Toll-like Receptor 7 Occurs upon Viral Entry. Immunity. 2011b;35:135–145. [PMC free article] [PubMed] [Google Scholar]
  • Kearney M, Maldarelli F, Shao W, Margolick JB, Daar ES, Mellors JW, Rao V, Coffin JM, Palmer S. Human immunodeficiency virus type 1 population genetics and adaptation in newly infected individuals. J Virol. 2009;83:2715–2727. [PMC free article] [PubMed] [Google Scholar]
  • Kennedy EM, Amie SM, Bambara RA, Kim B. Frequent incorporation of ribonucleotides during HIV-1 reverse transcription and their attenuated repair in macrophages. J Biol Chem. 2012;287:14280–14288. [PMC free article] [PubMed] [Google Scholar]
  • Kennedy EM, Daddacha W, Slater R, Gavegnano C, Fromentin E, Schinazi RF, Kim B. Abundant non-canonical dUTP found in primary human macrophages drives its frequent incorporation by HIV-1 reverse transcriptase. J Biol Chem. 2011;286:25047–25055. [PMC free article] [PubMed] [Google Scholar]
  • Kennedy EM, Gavegnano C, Nguyen L, Slater R, Lucas A, Fromentin E, Schinazi RF, Kim B. Ribonucleoside triphosphates as substrate of human immunodeficiency virus type 1 reverse transcriptase in human macrophages. J Biol Chem. 2010;285:39380–39391. [PMC free article] [PubMed] [Google Scholar]
  • Khaitan A, Unutmaz D. Revisiting immune exhaustion during HIV infection. Curr HIV/AIDS Rep. 2011;8:4–11. [PMC free article] [PubMed] [Google Scholar]
  • Kim B, Nguyen LA, Daddacha W, Hollenbaugh JA. Tight Interplay among SAMHD1 Protein Level, Cellular dNTP Levels, and HIV-1 Proviral DNA Synthesis Kinetics in Human Primary Monocyte-derived Macrophages. J Biol Chem. 2012;287:21570–21574. [PMC free article] [PubMed] [Google Scholar]
  • Kratky W, Reis e Sousa C, Oxenius A, Spörri R. Direct activation of antigen-presenting cells is required for CD8+ T-cell priming and tumor vaccination. Proceedings of the National Academy of Sciences. 2011;108:17414–17419. [PMC free article] [PubMed] [Google Scholar]
  • Laguette N, Sobhian B, Casartelli N, Ringeard M, Chable-Bessia C, Ségéral E, Yatim A, Emiliani S, Schwartz O, Benkirane M. SAMHD1 is the dendritic- and myeloid-cell-specific HIV-1 restriction factor counteracted by Vpx. Nature. 2011;474:654–657. [PMC free article] [PubMed] [Google Scholar]
  • Lahouassa H, Daddacha W, Hofmann H, Ayinde D, Logue EC, iuml c Dragin Lo, Bloch N, Maudet C, Bertrand M, Gramberg T, et al. SAMHD1 restricts the replication of human immunodeficiency virus type 1 by depleting the intracellular pool of deoxynucleoside triphosphates. Nat Immunol. 2012:1–7. [PMC free article] [PubMed] [Google Scholar]
  • Li Y, Chen R, Zhou Q, Xu Z, Li C, Wang S, Mao A, Zhang X, He W, Shu HB. LSm14A is a processing body-associated sensor of viral nucleic acids that initiates cellular antiviral response in the early phase of viral infection. Proceedings of the National Academy of Sciences. 2012;109:11770–11775. [PMC free article] [PubMed] [Google Scholar]
  • Liberatore RA, Bieniasz PD. Tetherin is a key effector of the antiretroviral activity of type I interferon in vitro and in vivo. Proceedings of the National Academy of Sciences. 2011;108:18097–18101. [PMC free article] [PubMed] [Google Scholar]
  • Lindell DM, Morris SB, White MP, Kallal LE, Lundy PK, Hamouda T, Baker JR, Lukacs NW. A novel inactivated intranasal respiratory syncytial virus vaccine promotes viral clearance without Th2 associated vaccine-enhanced disease. PLoS ONE. 2011;6:e21823. [PMC free article] [PubMed] [Google Scholar]
  • Longhi MP, Trumpfheller C, Idoyaga J, Caskey M, Matos I, Kluger C, Salazar AM, Colonna M, Steinman RM. Dendritic cells require a systemic type I interferon response to mature and induce CD4+ Th1 immunity with poly IC as adjuvant. J Exp Med. 2009;206:1589–1602. [PMC free article] [PubMed] [Google Scholar]
  • Lubong Sabado R, Kavanagh DG, Kaufmann DE, Fru K, Babcock E, Rosenberg E, Walker B, Lifson J, Bhardwaj N, Larsson M. In vitro priming recapitulates in vivo HIV-1 specific T cell responses, revealing rapid loss of virus reactive CD4 T cells in acute HIV-1 infection. PLoS ONE. 2009;4:e4256. [PMC free article] [PubMed] [Google Scholar]
  • Manel N, Hogstad B, Wang Y, Levy DE, Unutmaz D, Littman DR. A cryptic sensor for HIV-1 activates antiviral innate immunity in dendritic cells. Nature. 2010;467:214–217. [PMC free article] [PubMed] [Google Scholar]
  • Mangeot PE, Duperrier K, Nègre D, Boson B, Rigal D, Cosset FL, Darlix JL. High levels of transduction of human dendritic cells with optimized SIV vectors. Mol Ther. 2002;5:283–290. [PubMed] [Google Scholar]
  • McDonald D, Vodicka MA, Lucero G, Svitkina TM, Borisy GG, Emerman M, Hope TJ. Visualization of the intracellular behavior of HIV in living cells. J Cell Biol. 2002;159:441–452. [PMC free article] [PubMed] [Google Scholar]
  • McDonald D, Wu L, Bohks SM, Kewalramani VN, Unutmaz D, Hope TJ. Recruitment of HIV and its receptors to dendritic cell-T cell junctions. Science. 2003;300:1295–1297. [PubMed] [Google Scholar]
  • Moris A, Nobile C, Buseyne F, Porrot F, Abastado JP, Schwartz O. DC-SIGN promotes exogenous MHC-I-restricted HIV-1 antigen presentation. Blood. 2004;103:2648–2654. [PubMed] [Google Scholar]
  • Moris A, Pajot A, Blanchet F, Guivel-Benhassine F, Salcedo M, Schwartz O. Dendritic cells and HIV-specific CD4+ T cells: HIV antigen presentation, T-cell activation, and viral transfer. Blood. 2006;108:1643–1651. [PubMed] [Google Scholar]
  • Nabel GJ, Kwong PD, Mascola JR. Progress in the rational design of an AIDS vaccine. Philosophical Transactions of the Royal Society B: Biological Sciences. 2011;366:2759–2765. [PMC free article] [PubMed] [Google Scholar]
  • Peretz Y, He Z, Shi Y, Yassine-Diab B, Goulet J-P, Bordi R, Filali-Mouhim A, Loubert J-B, Far, El M, Dupuy FP, et al. CD160 and PD-1 Co-Expression on HIV-Specific CD8 T Cells Defines a Subset with Advanced Dysfunction. PLoS Pathog. 2012;8:e1002840. [PMC free article] [PubMed] [Google Scholar]
  • Perrino FW, Harvey S, Shaban NM, Hollis T. RNaseH2 mutants that cause Aicardi-Goutieres syndrome are active nucleases. J Mol Med. 2009;87:25–30. [PMC free article] [PubMed] [Google Scholar]
  • Pertel T, Hausmann S, Morger D, Züger S, Guerra J, Lascano J, Reinhard C, Santoni FA, Uchil PD, Chatel L, et al. TRIM5 is an innate immune sensor for the retrovirus capsid lattice. Nature. 2011a;472:361–365. [PMC free article] [PubMed] [Google Scholar]
  • Pertel T, Reinhard C, Luban J. Vpx rescues HIV-1 transduction of dendritic cells from the antiviral state established by type 1 interferon. Retrovirology. 2011b;8:49. [PMC free article] [PubMed] [Google Scholar]
  • Piguet V, Steinman RM. The interaction of HIV with dendritic cells: outcomes and pathways. Trends Immunol. 2007;28:503–510. [PMC free article] [PubMed] [Google Scholar]
  • Pope M, Betjes MG, Romani N, Hirmand H, Cameron PU, Hoffman L, Gezelter S, Schuler G, Steinman RM. Conjugates of dendritic cells and memory T lymphocytes from skin facilitate productive infection with HIV-1. Cell. 1994;78:389–398. [PubMed] [Google Scholar]
  • Powell RD, Holland PJ, Hollis T, Perrino FW. Aicardi-Goutieres Syndrome Gene and HIV-1 Restriction Factor SAMHD1 Is a dGTP-regulated Deoxynucleotide Triphosphohydrolase. J Biol Chem. 2011;286:43596–43600. [PMC free article] [PubMed] [Google Scholar]
  • Puryear WB, Yu X, Ramirez NP, Reinhard BM, Gummuluru S. HIV-1 incorporation of host-cell-derived glycosphingolipid GM3 allows for capture by mature dendritic cells. Proceedings of the National Academy of Sciences. 2012;109:7475–7480. [PMC free article] [PubMed] [Google Scholar]
  • Reis e Sousa C. Dendritic cells in a mature age. Nat Rev Immunol. 2006;6:476–483. [PubMed] [Google Scholar]
  • Rerks-Ngarm S, Pitisuttithum P, Nitayaphan S, Kaewkungwal J, Chiu J, Paris R, Premsri N, Namwat C, de Souza M, Adams E, et al. Vaccination with ALVAC and AIDSVAX to prevent HIV-1 infection in Thailand. N Engl J Med. 2009;361:2209–2220. [PubMed] [Google Scholar]
  • Rice GI, Bond J, Asipu A, Brunette RL, Manfield IW, Carr IM, Fuller JC, Jackson RM, Lamb T, Briggs TA, et al. Mutations involved in Aicardi-Goutières syndrome implicate SAMHD1 as regulator of the innate immune response. Nat Genet. 2009;41:829–832. [PMC free article] [PubMed] [Google Scholar]
  • Romagnani C, Chiesa, Della M, Kohler S, Moewes B, Radbruch A, Moretta L, Moretta A, Thiel A. Activation of human NK cells by plasmacytoid dendritic cells and its modulation by CD4+ T helper cells and CD4+ CD25hi T regulatory cells. Eur J Immunol. 2005;35:2452–2458. [PubMed] [Google Scholar]
  • Sabchareon A, Wallace D, Sirivichayakul C, Limkittikul K, Chanthavanich P, Suvannadabba S, Jiwariyavej V, Dulyachai W, Pengsaa K, Wartel TA, et al. Protective efficacy of the recombinant, live-attenuated, CYD tetravalent dengue vaccine in Thai schoolchildren: a randomised, controlled phase 2b trial. Lancet 2012 [PubMed] [Google Scholar]
  • Sawyer SL, Wu LI, Emerman M, Malik HS. Positive selection of primate TRIM5alpha identifies a critical species-specific retroviral restriction domain. Proc Natl Acad Sci USA. 2005;102:2832–2837. [PMC free article] [PubMed] [Google Scholar]
  • Sayah DM, Sokolskaja E, Berthoux L, Luban J. Cyclophilin A retrotransposition into TRIM5 explains owl monkey resistance to HIV-1. Nature. 2004;430:569–573. [PubMed] [Google Scholar]
  • Scheid JF, Mouquet H, Feldhahn N, Seaman MS, Velinzon K, Pietzsch J, Ott RG, Anthony RM, Zebroski H, Hurley A, et al. Broad diversity of neutralizing antibodies isolated from memory B cells in HIV-infected individuals. Nature. 2009;458:636–640. [PubMed] [Google Scholar]
  • Schröfelbauer B, Hakata Y, Landau NR. HIV-1 Vpr function is mediated by interaction with the damage-specific DNA-binding protein DDB1. Proc Natl Acad Sci USA. 2007;104:4130–4135. [PMC free article] [PubMed] [Google Scholar]
  • Schröfelbauer B, Yu Q, Zeitlin SG, Landau NR. Human immunodeficiency virus type 1 Vpr induces the degradation of the UNG and SMUG uracil-DNA glycosylases. J Virol. 2005;79:10978–10987. [PMC free article] [PubMed] [Google Scholar]
  • Shan L, Deng K, Shroff NS, Durand CM, Rabi SA, Yang HC, Zhang H, Margolick JB, Blankson JN, Siliciano RF. Stimulation of HIV-1-specific cytolytic T lymphocytes facilitates elimination of latent viral reservoir after virus reactivation. Immunity. 2012;36:491–501. [PMC free article] [PubMed] [Google Scholar]
  • Sharma S, Fitzgerald KA. Innate immune sensing of DNA. PLoS Pathog. 2011;7:e1001310. [PMC free article] [PubMed] [Google Scholar]
  • Sharma S, DeOliveira RB, Kalantari P, Parroche P, Goutagny N, Jiang Z, Chan J, Bartholomeu DC, Lauw F, Hall JP, et al. Innate immune recognition of an AT-rich stem-loop DNA motif in the Plasmodium falciparum genome. Immunity. 2011;35:194–207. [PMC free article] [PubMed] [Google Scholar]
  • Sharova N, Wu Y, Zhu X, Stranska R, Kaushik R, Sharkey M, Stevenson M. Primate lentiviral Vpx commandeers DDB1 to counteract a macrophage restriction. PLoS Pathog. 2008;4:e1000057. [PMC free article] [PubMed] [Google Scholar]
  • Smed-Sörensen A, Loré K, Vasudevan J, Louder MK, Andersson J, Mascola JR, Spetz AL, Koup RA. Differential susceptibility to human immunodeficiency virus type 1 infection of myeloid and plasmacytoid dendritic cells. J Virol. 2005;79:8861–8869. [PMC free article] [PubMed] [Google Scholar]
  • Smed-Sörensen A, Loré K, Walther-Jallow L, Andersson J, Spetz AL. HIV-1-infected dendritic cells up-regulate cell surface markers but fail to produce IL-12 p70 in response to CD40 ligand stimulation. Blood. 2004;104:2810–2817. [PubMed] [Google Scholar]
  • Smith DM, Richman DD, Little SJ. HIV superinfection. J Infect Dis. 2005;192:438–444. [PubMed] [Google Scholar]
  • Sokolskaja E, Luban J. Cyclophilin, TRIM5, and innate immunity to HIV-1. Curr Opin Microbiol. 2006;9:404–408. [PubMed] [Google Scholar]
  • Sokolskaja E, Berthoux L, Luban J. Cyclophilin A and TRIM5alpha independently regulate human immunodeficiency virus type 1 infectivity in human cells. J Virol. 2006;80:2855–2862. [PMC free article] [PubMed] [Google Scholar]
  • Solis M, Nakhaei P, Jalalirad M, Lacoste J, Douville R, Arguello M, Zhao T, Laughrea M, Wainberg MA, Hiscott J. RIG-I-mediated antiviral signaling is inhibited in HIV-1 infection by a protease-mediated sequestration of RIG-I. J Virol. 2011;85:1224–1236. [PMC free article] [PubMed] [Google Scholar]
  • Spörri R, Reis e Sousa C. Inflammatory mediators are insufficient for full dendritic cell activation and promote expansion of CD4+ T cell populations lacking helper function. Nat Immunol. 2005;6:163–170. [PubMed] [Google Scholar]
  • Srivastava S, Swanson SK, Manel N, Florens L, Washburn MP, Skowronski J. Lentiviral Vpx accessory factor targets VprBP/DCAF1 substrate adaptor for cullin 4 E3 ubiquitin ligase to enable macrophage infection. PLoS Pathog. 2008;4:e1000059. [PMC free article] [PubMed] [Google Scholar]
  • Steinman RM, Idoyaga J. Features of the dendritic cell lineage. Immunol Rev. 2010;234:5–17. [PubMed] [Google Scholar]
  • Stetson DB, Ko JS, Heidmann T, Medzhitov R. Trex1 prevents cell-intrinsic initiation of autoimmunity. Cell. 2008;134:587–598. [PMC free article] [PubMed] [Google Scholar]
  • Stremlau M, Owens CM, Perron MJ, Kiessling M, Autissier P, Sodroski J. The cytoplasmic body component TRIM5alpha restricts HIV-1 infection in Old World monkeys. Nature. 2004;427:848–853. [PubMed] [Google Scholar]
  • Stremlau M, Perron M, Lee M, Li Y, Song B, Javanbakht H, Diaz-Griffero F, Anderson DJ, Sundquist WI, Sodroski J. Specific recognition and accelerated uncoating of retroviral capsids by the TRIM5alpha restriction factor. Proc Natl Acad Sci USA. 2006;103:5514–5519. [PMC free article] [PubMed] [Google Scholar]
  • Stremlau M, Perron M, Welikala S, Sodroski J. Species-specific variation in the B30.2(SPRY) domain of TRIM5alpha determines the potency of human immunodeficiency virus restriction. J Virol. 2005;79:3139–3145. [PMC free article] [PubMed] [Google Scholar]
  • Sunseri N, O’Brien M, Bhardwaj N, Landau NR. Human immunodeficiency virus type 1 modified to package Simian immunodeficiency virus Vpx efficiently infects macrophages and dendritic cells. J Virol. 2011;85:6263–6274. [PMC free article] [PubMed] [Google Scholar]
  • Tareen SU, Emerman M. Human Trim5α has additional activities that are uncoupled from retroviral capsid recognition. Virology. 2011;409:113–120. [PMC free article] [PubMed] [Google Scholar]
  • Tatsuo H, Ono N, Tanaka K, Yanagi Y. SLAM (CDw150) is a cellular receptor for measles virus. Nature. 2000;406:893–897. [PubMed] [Google Scholar]
  • Trautmann L, Janbazian L, Chomont N, Said EA, Gimmig S, Bessette B, Boulassel MR, Delwart E, Sepulveda H, Balderas RS, et al. Upregulation of PD-1 expression on HIV-specific CD8+ T cells leads to reversible immune dysfunction. Nat Med. 2006;12:1198–1202. [PubMed] [Google Scholar]
  • Unterholzner L, Keating SE, Baran M, Horan KA, Jensen SB, Sharma S, Sirois CM, Jin T, Latz E, Xiao TS, et al. IFI16 is an innate immune sensor for intracellular DNA. Nat Immunol. 2010;11:997–1004. [PMC free article] [PubMed] [Google Scholar]
  • Upton JW, Kaiser WJ, Mocarski ES. DAI/ZBP1/DLM-1 complexes with RIP3 to mediate virus-induced programmed necrosis that is targeted by murine cytomegalovirus vIRA. Cell Host Microbe. 2012;11:290–297. [PMC free article] [PubMed] [Google Scholar]
  • Vignuzzi M, Stone JK, Arnold JJ, Cameron CE, Andino R. Quasispecies diversity determines pathogenesis through cooperative interactions in a viral population. Nature. 2006;439:344–348. [PMC free article] [PubMed] [Google Scholar]
  • Virgin HW, Walker BD. Immunology and the elusive AIDS vaccine. Nature. 2010;464:224–231. [PubMed] [Google Scholar]
  • Welliver RC. Review of epidemiology and clinical risk factors for severe respiratory syncytial virus (RSV) infection. J Pediatr. 2003;143:S112–S117. [PubMed] [Google Scholar]
  • Wherry EJ, Ha SJ, Kaech SM, Haining WN, Sarkar S, Kalia V, Subramaniam S, Blattman JN, Barber DL, Ahmed R. Molecular signature of CD8+ T cell exhaustion during chronic viral infection. Immunity. 2007;27:670–684. [PubMed] [Google Scholar]
  • Yan N, O’Day E, Wheeler LA, Engelman A, Lieberman J. HIV DNA is heavily uracilated, which protects it from autointegration. Proceedings of the National Academy of Sciences. 2011;108:9244–9249. [PMC free article] [PubMed] [Google Scholar]
  • Yan N, Regalado-Magdos AD, Stiggelbout B, Lee-Kirsch MA, Lieberman J. The cytosolic exonuclease TREX1 inhibits the innate immune response to human immunodeficiency virus type 1. Nat Immunol. 2010;11:1005–1013. [PMC free article] [PubMed] [Google Scholar]
  • Yoon C, Johnston SC, Tang J, Stahl M, Tobin JF, Somers WS. Charged residues dominate a unique interlocking topography in the heterodimeric cytokine interleukin-12. Embo J. 2000;19:3530–3541. [PMC free article] [PubMed] [Google Scholar]
  • Zhang Z, Yuan B, Bao M, Lu N, Kim T, Liu YJ. The helicase DDX41 senses intracellular DNA mediated by the adaptor STING in dendritic cells. Nat Immunol. 2011;12:959–965. [PMC free article] [PubMed] [Google Scholar]
  • Zhao G, Ke D, Vu T, Ahn J, Shah VB, Yang R, Aiken C, Charlton LM, Gronenborn AM, Zhang P. Rhesus TRIM5α disrupts the HIV-1 capsid at the inter-hexamer interfaces. PLoS Pathog. 2011;7:e1002009. [PMC free article] [PubMed] [Google Scholar]
  • Zhou T, Xu L, Dey B, Hessell AJ, Van Ryk D, Xiang SH, Yang X, Zhang MY, Zwick MB, Arthos J, et al. Structural definition of a conserved neutralization epitope on HIV-1 gp120. Nature. 2007;445:732–737. [PMC free article] [PubMed] [Google Scholar]
-