HIV/AIDS has long been at the forefront of the development of gene- and cell-based therapies. Although conventional gene therapy approaches typically involve the addition of anti-HIV genes to cells using semirandomly integrating viral vectors, newer genome editing technologies based on engineered nucleases are now allowing more precise genetic manipulations. The possible outcomes of genome editing include gene disruption, which has been most notably applied to the CCR5 coreceptor gene, or the introduction of small mutations or larger whole gene cassette insertions at a targeted locus. Disruption of CCR5 using zinc finger nucleases was the first-in-human application of genome editing and remains the most clinically advanced platform, with 7 completed or ongoing clinical trials in T cells and hematopoietic stem/progenitor cells (HSPCs). Here we review the laboratory and clinical findings of CCR5 editing in T cells and HSPCs for HIV therapy and summarize other promising genome editing approaches for future clinical development. In particular, recent advances in the delivery of genome editing reagents and the demonstration of highly efficient homology-directed editing in both T cells and HSPCs are expected to spur the development of even more sophisticated applications of this technology for HIV therapy.

HIV-1 infection presents a unique set of challenges for the development of effective therapies. During infection, the virus inserts a copy of itself into the genome of a cell and thereby becomes a permanent part of the cell’s genetic material. On occasion, expression from such integrated viruses is suppressed, leading to a latent viral state that is not impacted by the antiretroviral drug therapies (ARTs) that target replicating viruses. Although rare,1  such latently infected cells comprise a significant reservoir in the body with a long half-life. Moreover, latent virus can be reactivated in response to changes in a cell’s activation status, which leads to a rapid rebound of viremia if ART is discontinued.2-4  As a result, HIV management requires life-long ART, which does not ever cure people.

Adherence to a daily ART regimen can be challenging. Compliance is impacted by accessibility, drug-related side effects, lack of health literacy, and social and emotional challenges. Consequently, a significant portion of patients on ART do not achieve and maintain full virologic control, and even in a developed country with highly accessible ART such as the United States, only a third of the estimated 1.2 million people currently infected with HIV-1 achieve optimal ART.5  HIV infection, even when fully suppressed, can continue to impact immune function, in particular as a result of the decimation of CD4+ T cells that occurs during the acute initial phase of infection. Indeed, some patients never recover normal levels of CD4+ T cells, even when ART provides effective virologic control.6 

The limitations of ART and the similarities between HIV infection and a more traditional genetic disease made HIV/AIDS an early poster child for the development of experimental genetic-, cellular-, and immune-based therapies.7-9  Despite the recognized complexity of such approaches, the possibilities of one-shot treatments that could both restrict viral replication and restore immune function are spurring development. To date, most gene therapy approaches for HIV have involved the addition of anti-HIV genes to protect CD4+ T cells from infection. By reducing the number of susceptible cells, such strategies are expected to reduce overall viremia, while the selective survival of the engineered cells would increase the effectiveness of the treatment over time. In addition, the direct protection of CD4+ T cells could restore immune function, allowing the body to fight back against both the symptoms of the infection and the virus itself. A virtuous cycle is thus envisioned.

Several anti-HIV genes have already been tested in this way in clinical trials, including dominant negative viral proteins (Rev M10) and peptides (C46), viral RNA decoys (TAR and RRE), and RNA-based methods to suppress either host or viral genes using ribozymes, antisense RNA, and RNA interference technologies.10-16  The candidate anti-HIV factors are delivered ex vivo to either CD4+ T cells or their in vivo precursors: the hematopoietic stem and progenitor cells (HSPCs). To achieve permanent gene expression, integrating retroviral and lentiviral vectors are typically used, taking advantage of viral machineries that have evolved to permanently integrate into the genomes of hematopoietic cells. On the down side, these vectors can suffer from lack of long-term gene expression and genotoxic risk due to random genomic integration.17,18  Clinical trials to date have established a portfolio of candidate anti-HIV genes, demonstrated safety, and created an appetite for gene therapy, but stopped short of demonstrating efficacy.

Genome editing has emerged recently as a more precise way to engineer cells that could also be applied to HIV/AIDS. At its heart, the technology uses engineered nucleases to introduce DNA double-strand breaks (DSBs) at a targeted locus, whose subsequent repair is exploited to achieve different outcomes. If the nonhomologous end joining (NHEJ) repair pathway is accessed, the end result can be random insertions/deletions (indels) at the break site that result in gene disruption. Alternatively, more precise repair pathways based on homologous recombination can be hijacked to copy information from an introduced DNA homology template. Such homology-directed repair (HDR) can promote a specific gene editing/mutation event or allow the site-specific addition of larger gene cassettes at the break site. The initial DSB generation can be achieved using different engineered nuclease platforms, including homing endonucleases, zinc finger nucleases (ZFNs), transcription activator–like effector nucleases (TALENs), and Clustered regularly interspaced short palindromic repeat (CRISPR)/Cas9.19,20  As described below, all 3 of the possible outcomes of DSB repair are being developed for HIV-1 treatment (Figure 1).

Figure 1

Potential applications of genome editing in anti-HIV therapy. Site-specific DNA breaks created by engineered nucleases can be repaired by error-prone NHEJ or, if a homologous DNA repair template is also present, by more precise HDR pathways. NHEJ frequently results in indels, allowing disruption of genes such as the HIV-1 coreceptors CCR5 and CXCR4 or integrated HIV-1 genomes. HDR could be used to introduce small mutations into host restriction factors to restore anti-HIV activities or into host dependency factors to limit their use by the virus. Alternatively, HDR could be used to site-specifically insert anti-HIV genes at the site of the DNA break, including at a disrupted CCR5 locus.

Figure 1

Potential applications of genome editing in anti-HIV therapy. Site-specific DNA breaks created by engineered nucleases can be repaired by error-prone NHEJ or, if a homologous DNA repair template is also present, by more precise HDR pathways. NHEJ frequently results in indels, allowing disruption of genes such as the HIV-1 coreceptors CCR5 and CXCR4 or integrated HIV-1 genomes. HDR could be used to introduce small mutations into host restriction factors to restore anti-HIV activities or into host dependency factors to limit their use by the virus. Alternatively, HDR could be used to site-specifically insert anti-HIV genes at the site of the DNA break, including at a disrupted CCR5 locus.

Close modal

It is no accident that the first application of targeted nucleases in humans was an anti-HIV strategy based on preventing expression of the CCR5 gene. Achieving gene disruption is far simpler than gene editing or insertion, because it relies only on the activity of the nuclease and does not require the additional delivery of a homology template. In addition, NHEJ is active throughout the cell cycle, whereas HDR is mainly restricted to S and G2, when repair templates are available in the form of sister chromatids.21,22  Finally, CCR5 represents a uniquely attractive disruption target, being a nonessential human gene whose absence provides potent anti-HIV protection.

CCR5 is the entry coreceptor used by the majority of HIV strains and in particular by the transmitting and early infecting strains.23  The CCR5Δ32 allele contains a 32-bp deletion that results in a truncated protein that is not expressed on the cell surface. The allele confers protection against HIV-1 infection without adverse health consequences in homozygotes, which constitute ∼1% of the Caucasian population, and HIV-infected heterozygotes exhibit delayed disease progression.24-26 

The potential benefit of a CCR5 targeted gene therapy was strikingly demonstrated in 2009, in the first and only reported case of an HIV cure. The “Berlin patient” received an allogeneic HSPC transplant from a CCR5Δ32 donor during treatment of acute myeloid leukemia and has remained HIV-1 free without ART since then.27,28  Although he received significant conditioning as part of his treatment and developed graft-versus-host-disease, both of which likely reduced his HIV-1 reservoir, the CCR5Δ32 genotype of his donor is considered an essential component of his cure, because other attempts using HSPC transplants from CCR5 wild-type donors have not been curative.29,30 

The extraordinary finding of a single individual cured of HIV-1 galvanized efforts to use gene therapy to create CCR5-negative autologous T cells or HSPCs in HIV-infected patients, with the goal of providing the same curative benefits without the challenges of allogeneic transplantation. At the same time, genome engineering emerged as a preferred method to achieve complete and permanent CCR5 depletion,31  which did not suffer from the incomplete or nonpermanent effects of protein- and RNA-based strategies.32-36  However, the clinical implementation of genome engineering required considerable development efforts to identify protocols that could deliver engineered nucleases to target cells and maximize bi-allelic disruption, while limiting off-target effects at nontargeted sites. This entirely new class of gene therapy technologies also required the development of appropriate in vitro and in vivo model systems to allow the necessary and rigorous preclinical safety studies.37  Much like it has in the past for gene therapy, HIV disease has been a pioneer in these endeavors.

The first clinical trials to evaluate genome engineered CCR5-negative cells used ZFNs and built on prior experiences in T-cell adoptive transfer. A ZFN pair was identified that introduced a DSB at approximately nucleotide 160 of the CCR5 open-reading frame, corresponding to a site in the first of 7 transmembrane domains in the protein.38  The most frequent outcome of treatment with this ZFN pair is a 5 nucleotide addition, which accounts for ∼25% of all modified alleles.38,39  This pentamer duplication introduces 2 in-frame stop codons that block expression and also serves as a convenient genetic marker for estimating the overall frequency of CCR5 modification (Figure 2).

Figure 2

ZFN-mediated disruption of the CCR5 open-reading frame. A ZFN pair that binds CCR5 sequences (left and right, boxed) results in a DNA break whose repair most frequently results in a 5-bp duplication (arrowed) that introduces premature stop codons (*) into the open-reading frame.

Figure 2

ZFN-mediated disruption of the CCR5 open-reading frame. A ZFN pair that binds CCR5 sequences (left and right, boxed) results in a DNA break whose repair most frequently results in a 5-bp duplication (arrowed) that introduces premature stop codons (*) into the open-reading frame.

Close modal

The anti-HIV efficacy of the ZFN pair was first demonstrated in preclinical studies using primary CD4+ T cells.38  A chimeric Ad5/F35 adenoviral vector was used to deliver the ZFN pair and was found capable of disrupting up to 40% to 60% of total CCR5 alleles. Furthermore, 33% of modified cells were disrupted at both alleles and therefore were effectively CCR5 null. In a mouse xenotransplantation model, HIV-1 challenge was seen to increase the proportion of modified CCR5 alleles threefold, confirming the expected survival advantage for the gene-edited cells. Mice receiving CCR5 ZFN-treated cells also better preserved their human CD4+ T cells and had lower viremia following HIV-1 challenge.38  Subsequently, Ad5/F35 delivery of this ZFN pair to CD3/CD28-stimulated CD4+ T cells was adapted for clinical scale, allowing the production of >1010CCR5-edited cells and paving the way for phase 1 clinical studies.40 

The first-in-human genome editing trial to treat HIV (#NCT00842634) was initiated in 2009, primarily to evaluate the safety of modifying autologous CD4+ T cells in HIV-1–infected individuals41,42  (Table 1). Twelve patients on ART, with undetectable viral loads, were enrolled in 2 cohorts, based on whether their CD4+ T-cell counts were >450/mm3 (cohort 1) or at less optimal levels of 200 to 500/mm3 (cohort 2). Each participant received a single infusion of 5 to 10 billion ZFN-modified autologous CD4+ T cells. The infusions were well tolerated, with only 1 individual experiencing a minor infusion-related adverse event.

Table 1

Clinical trials of HIV-infected individuals using ZFN-modified autologous T cells and HSPCs

Clinical trialStatusCohorts/study populationsCell typeZFN delivery methodConditioning regimen
NCT00842634 phase 1 2009-2013 Completed41  On ART, aviremic, CD4 counts >450 CD4 T cells Adenovirus None 
On ART, aviremic, CD4 counts 200-500, ATI 
NCT01044654 phase 1 2009-2014 Completed On ART, aviremic, CD4 counts 200-500 CD4 T cells Adenovirus None 
On ART, aviremic, CD4 counts >500, and CCR5Δ32 heterozygotes, ATI 
NCT01252641 phase 1/2 2010-2015 Completed Not on ART, CD4 counts >500 CD4 T cells Adenovirus None 
NCT01543152 phase 1 2011-2016 Recruiting On ART, aviremic, CD4 counts >500, ATI CD4 T cells, and CD4/CD8 T cells Adenovirus Cytoxan (dose escalation) 
NCT02225665 phase 1/2 2014-2018 Active, not recruiting On ART, aviremic, CD4 counts >500, ATI CD4/CD8 T cells mRNA Cytoxan (1 g/m2
NCT02388594 phase 1 2015-2017 Recruiting On ART, aviremic, CD4 counts >450, ATI CD4 T cells mRNA Cytoxan (1 g/m2
NCT02500849 phase 1 2015-2018 Recruiting On ART, aviremic, CD4 counts 200-500, ATI CD34 HSPC mRNA Busulfan (dose escalation) 
Clinical trialStatusCohorts/study populationsCell typeZFN delivery methodConditioning regimen
NCT00842634 phase 1 2009-2013 Completed41  On ART, aviremic, CD4 counts >450 CD4 T cells Adenovirus None 
On ART, aviremic, CD4 counts 200-500, ATI 
NCT01044654 phase 1 2009-2014 Completed On ART, aviremic, CD4 counts 200-500 CD4 T cells Adenovirus None 
On ART, aviremic, CD4 counts >500, and CCR5Δ32 heterozygotes, ATI 
NCT01252641 phase 1/2 2010-2015 Completed Not on ART, CD4 counts >500 CD4 T cells Adenovirus None 
NCT01543152 phase 1 2011-2016 Recruiting On ART, aviremic, CD4 counts >500, ATI CD4 T cells, and CD4/CD8 T cells Adenovirus Cytoxan (dose escalation) 
NCT02225665 phase 1/2 2014-2018 Active, not recruiting On ART, aviremic, CD4 counts >500, ATI CD4/CD8 T cells mRNA Cytoxan (1 g/m2
NCT02388594 phase 1 2015-2017 Recruiting On ART, aviremic, CD4 counts >450, ATI CD4 T cells mRNA Cytoxan (1 g/m2
NCT02500849 phase 1 2015-2018 Recruiting On ART, aviremic, CD4 counts 200-500, ATI CD34 HSPC mRNA Busulfan (dose escalation) 

Information obtained from website Clinicaltrials.gov and publically reported by sponsors.

In terms of safety, the CCR5 ZFN-modified cells displayed normal characteristics, including engraftment in all patients, persistence for ≥42 months after infusion, and normal trafficking to the rectal mucosa. The trial also attempted to discern any anti-HIV effects. Four weeks after infusion, cohort 1 patients underwent a 12-week analytical treatment interruption (ATI), during which ART was stopped. This usually results in rapid viral rebound within 2 to 4 weeks and the establishment of a plasma viral load set point that is similar to historical levels in each patient before ART initiation. Changes in either the time to rebound or the viral load set point could indicate an anti-HIV effect. In addition, allowing a limited period of HIV replication during ART cessation was hypothesized to enrich for the CCR5-negative, HIV-resistant cells.

In the 4 patients that completed ATI, HIV rebounded and reached a peak viremia during ART withdrawal. Although CD4+ T cell counts decreased in all patients during this period of viremia, the mean rate of decline of the CCR5-modified CD4+ T cells was slower than that of the unmodified cells (−1.81 vs −7.25 cells/mm3 per day), suggesting a protective effect of the CCR5 modifications. Notably, 1 patient had a particularly unusual response, showing both delayed viral rebound (6 weeks into ATI) and a peak viremia that was lower than the patient’s historical set point. Intriguingly, this was followed by a decrease in plasma HIV levels to below the limit of detection, before ART was resumed as planned. Subsequently, this patient was identified as being heterozygous for CCR5Δ32, leading to the hypothesis that this “halfway there” genotype had amplified the effects of the ZFN treatment.

Since this initial demonstration of clinical safety, subsequent trials have been designed to further optimize the treatment (Table 1). The parameters being evaluated include varying the input dose of cells, using multiple infusions of cells, including ZFN-modified CD8+ T cells in the graft, assessing the ability of Cytoxan to transiently reduce T-cell numbers and thereby improve engraftment of the infused T cells, and switching from adenoviral vector delivery of the ZFNs to using mRNA electroporation.

The possible impact of the heterozygous CCR5Δ32 genotype is also being assessed in a cohort of 10 patients in trial #NCT01044654. These patients underwent a 16-week ATI 2 months after infusion, during which time 3 patients were able to reduce and maintain viral loads at low (<1000 copies/mL) or undetectable levels. It has also been reported that 1 of the 3 patients maintained a viral load of <500 copies/mL, with CD4+ T-cell counts of >1000 cells/mL, for >1 year without ART.43,44  Long-term follow-up will determine whether a “functional cure” has been achieved in this patient.

A major concern for any clinical application of engineered nucleases is the potential for off-target genome modifications, reviewed elsewhere in this series. For the CCR5 ZFN pair in clinical development, a disruption rate of 5.39% has been observed in CD4+ T cells at the top predicted off-target site, the highly related CCR2 gene, under conditions that gave 36% on-target CCR5 disruption.38  The prediction, assessment, and mitigation of off-target activities have undergone major improvements in the last few years, so that if such off-target disruption at CCR2 proved to be problematic, it could be engineered around by selecting and developing a different ZFN pair.45,46  Of note, CCR2 disruption is predicted to be well tolerated and may even confer an additional anti–HIV-1 benefit, because a naturally occurring CCR2 mutant allele is associated with delayed progression to AIDS.47 

Engineering HSPCs instead of CD4+ T cells has the potential to provide a long-lasting source of modified cells and to additionally protect the CD4+ myeloid cells that are also susceptible to HIV-1. We previously demonstrated that the same CCR5 ZFN pair used in the T-cell trials could effectively modify CD34+ HSPCs, isolated from cord blood, fetal liver, or mobilized peripheral blood.39,48,49  Transplantation of ZFN-edited HSPCs into NSG mice was used to evaluate the safety of the treatment, with readouts in the ability of the treated cells to engraft and differentiate in the same manner as unmodified cells and to give rise to gene modified CD4+ T-cell progeny.50  Evidence of efficacy was also demonstrated by the selective preservation of gene-modified cells during HIV infection and by the suppression of viremia to levels below the limit of detection in mouse plasma.39,51 

The development of effective and clinically appropriate nuclease delivery methods for HSPCs has been a major challenge. Initial experiments using electroporation of plasmid DNA, although simple, ran into dose-limiting toxicities at higher amounts. Evaluation of Ad5/F35-mediated delivery into adult mobilized HSPCs revealed that this vector is less efficient in HSPCs than in CD4+ T cells, achieving only 5% CCR5 disruption, although frequencies could be increased by treatment with protein kinase C activators.48  Subsequently, we and others identified mRNA electroporation as an efficient and nontoxic method for ZFN introduction into HSPCs, achieving 30% to 50% CCR5 disruption with little toxicity.49,52,53  This method has now been adapted for large-scale clinical use in a recently initiated clinical trial (#NCT02500849) (Table 1).

Because HSPC gene therapy is often limited by difficulties in the ex vivo culture and expansion of HSPCs, there is also interest in the idea of modifying patient-specific induced pluripotent stem cells (iPSCs), which could then be reprogrammed into HSPCs.54  Clonal selection, differentiation, and expansion would allow 100% modification efficiencies to be achieved with homogeneity of the genome editing outcome while facilitating comprehensive testing for off-target effects. Several groups have reported making CCR5-disrupted CD34+ cells at small-scale from iPSCs using ZFNs, TALENs, and CRISPR/Cas9, and facilitated by HDR-mediated insertion of a selection cassette into CCR5.55,56  Further advances in reprogramming technologies may allow this to become a viable clinical alternative to modifying autologous HSPCs.

Beyond ZFNs, other nuclease platforms are also being developed to disrupt CCR5 in T cells and HSPCs, including TALENs,57  megaTALs,58  and CRISPR/Cas9.59,60  One potential advantage of CRISPR/Cas9 is that multiple guide RNAs can be used simultaneously to potentiate the magnitude of gene disruption or to induce specific deletions that are larger than the typical indels generated by NHEJ.59  However such multicomponent systems (requiring nuclease and guide RNAs) may introduce complexity into the clinical development process, and clinically suitable methods to deliver CRISPR/Cas9 to HSPCs in particular are still under development.61 

Although early infecting HIV-1 strains use CCR5 as the coreceptor, CXCR4-tropic viruses emerge in nearly half of patients in the later stages of infection.23,62  Disruption of CXCR4 is therefore also being explored as a strategy for patients who harbor CXCR4-tropic HIV-1. However, because CXCR4 plays an essential role in hematopoietic stem cell homing and retention in the bone marrow, this approach is limited to T cells.63  Two reports have described the ability of CXCR4-targeted ZFNs to protect human CD4+ T cells in humanized mice infected with CXCR4-tropic HIV-164,65  while the simultaneous disruption of both CCR5 and CXCR4 by ZFNs conferred protection against both tropisms of virus.66 

HDR-mediated gene editing and addition for anti-HIV therapies

More sophisticated application of genome editing will be achieved through exploiting HDR pathways. Compared with standard lentiviral vector approaches, HDR-mediated insertion of an anti-HIV gene offers the potential to provide greater control over the location, copy number, and expression profile of the added transgene. Those anti-HIV genes already evaluated in clinical trials remain candidates for this more precise method of gene addition.10-16  Moreover, by inserting such cassettes at a disrupted CCR5 locus, HIV resistance factors could be “stacked”67  and also provide protection to even mono-allelically modified cells. Beyond anti-HIV genes, genome editing technologies are also being developed to precisely insert HIV-specific chimeric antigen receptors into T cells and thereby exploit recent advances in immunotherapy to kill HIV-infected cells.58 

The capability of genome engineering to make small changes to genes could also be exploited to engineer alleles of human genes that are associated with better virologic control. Two broad classes of host genes could be considered as candidates: restriction factors and dependency factors.

Cellular restriction factors are host genes with broad antiviral properties and are often interferon inducible.68  They are also characterized by high evolution rates, reflecting the ongoing battle between host factors and the pathogens they target. Restriction factors with activity against HIV-1 include TRIM5α, tetherin, and APOBEC3G, but the evolutionary adaptations of HIV-1 to its human host mean that the human versions no longer function against this virus. In contrast, orthologs from nonhuman primates frequently retain anti-HIV activity, and comparisons between human and primate genes have identified single or small numbers of amino acids that could be edited into the human genes to restore anti-HIV activity.69-71 

In contrast, dependency factors72-74  are the group of cellular factors that are essential for HIV-1 replication and that are exploited by this viral parasite. Together with information from genome-wide association studies, which can identify genetic variants associated with greater HIV control,75  it may be possible to select candidate alleles from naturally occurring human variants that could also be engineered to inhibit HIV-1. The challenge will always be to find mutations that successfully prevent HIV from exploiting a host cell gene while not adversely impacting its function in noninfected cells.

Although these applications of genome editing suggest exciting new possibilities for HIV therapies, a major limitation to their development has been the low HDR editing efficiency that has been achieved in quiescent cell types such as mature T cells and HSPCs. One reason is that HDR is mainly active during the S and G2 phases of the cell cycle, but promoting cell cycling can adversely affect the function of quiescent cells. In addition, homology templates need to be present as DNA, which can cause cytotoxicity due to activation of DNA sensing pathways. Progress is now being made, as using integrase-defective lentiviral vectors as homology templates allowed gene editing rates of up to 5% at the IL2RG, CCR5, and AAVS1 loci in CD4+ T cells76  and at IL2RG in CD34+ HSCPs.53,76-78  More recently, we and others have described the use of adeno-associated virus (AAV) serotype 6 vectors in combination with electroporation of nuclease mRNA to achieved HDR-mediated genome editing. This method can be applied to both primary T cells and HSPCs, with little toxicity, and achieving editing rates of 8% to 60% at CCR5 alleles in primary T cells58,61  and 15% to 40% at the CCR5, IL2RG, HBB, and AAVS1 loci in mobilized blood HSPCs.49,58  Importantly, we also demonstrated that this approach edits the most primitive CD90+ cells in the bulk CD34+ population and can support serial transplantation of immune-deficient mice.49  Future methods to improve outcomes may come from manipulation of culture conditions to preserve gene-modified primitive HSPCs,53  or by the knocking-in of a selection marker, to allow further in vivo selection for successfully modified cells.79,80 

Disrupting integrated HIV-1 proviral genome

A major limitation of ART and many anti-HIV gene therapies that target different stages of the viral life cycle is that, although these approaches can prevent spread of infection, they do not eliminate integrated HIV-1 genomes. Because an integrated genome is to all intents and purposes a gene in the cell that carries it, the potential exists to use the capabilities of engineered nucleases to disrupt these unwanted genetic passengers. Moreover, such applications could theoretically also target latent genomes and thereby reduce the size of the HIV-1 reservoir that persists despite ART.

The first use of engineered nucleases to excise integrated HIV-1 genomes was reported by Sarkar et al, who evolved a Cre recombinase (Tre) to recognize a sequence within the HIV-1 long terminal repeat that had 50% similarity to the natural Cre recognition site, LoxP.81  This approach takes advantage of the ability of Cre to remove the intervening DNA sequence between two flanking LoxP sites. More recently, the same group has evolved a more broadly acting enzyme, named universal Tre (uTre), that targets an long terminal repeat sequence conserved across multiple subtypes of HIV-1.82  More conventional anti-HIV nucleases have also been described based on homing endonuclease, ZFNs, TALENs, and CRISPR/Cas9, targeting different regions of the HIV-1 genome, and with a demonstrated ability to reduce HIV-1 content in various cell lines.83-88 

Although these studies provide the initial proof of principle that gene disruption could be used to disrupt or excise an integrated HIV-1 genome, significant challenges will exist in the clinical translation of these approaches. No efficient method yet exists to enable in vivo delivery of nucleases to HIV target cells, and delivery to the subset of cells that comprise the latent reservoir will be especially problematic, due both to the location of these cells in inaccessible organ niches and the lack of identifying surface markers of viral infection. Alternatively, HIV-targeted nucleases could be delivered to bulk CD4+ T cells or HSPCs and be placed under inducible expression using an HIV-1 Tat-responsive promoter, as was reported in the case of the Tre recombinase.81  The recent reports of CD4-targeted lentiviral and AAV vectors suggest strategies toward targeted in vivo transduction of bulk CD4+ T cells.89,90 

HIV/AIDS provides a uniquely suitable disease for the development of applications of genome editing. ZFNs are currently the most clinically advanced engineered nuclease system being used, with 7 phase 1/2 clinical trials evaluating CCR5 disruption in T cells and HSPCs. In the future, engineered nucleases could also be used to disrupt the alternative HIV-1 coreceptor, CXCR4, or to target integrated HIV genomes directly. Moreover, recent demonstrations that highly efficient gene editing or addition is now possible in T cells and HSPCs through the use of AAV vectors to deliver homology templates should allow even more sophisticated applications of genome editing against this important human pathogen.

This work was supported by National Institutes of Health, National Heart, Lung, and Blood Institute grant HL129902 and California Institute for Regenerative Medicine grant RT3-07848.

Contribution: C.X.W. and P.M.C. wrote the manuscript.

Conflict-of-interest disclosure: The authors declare no competing financial interests.

Correspondence: Paula Cannon, Department of Molecular Microbiology & Immunology, Keck School of Medicine, University of Southern California, 2011 Zonal Ave, HMR413, Los Angeles, CA 90033; e-mail: pcannon@usc.edu.

1
Chun
 
TW
Carruth
 
L
Finzi
 
D
, et al. 
Quantification of latent tissue reservoirs and total body viral load in HIV-1 infection.
Nature
1997
, vol. 
387
 
6629
(pg. 
183
-
188
)
2
Finzi
 
D
Hermankova
 
M
Pierson
 
T
, et al. 
Identification of a reservoir for HIV-1 in patients on highly active antiretroviral therapy.
Science
1997
, vol. 
278
 
5341
(pg. 
1295
-
1300
)
3
Finzi
 
D
Blankson
 
J
Siliciano
 
JD
, et al. 
Latent infection of CD4+ T cells provides a mechanism for lifelong persistence of HIV-1, even in patients on effective combination therapy.
Nat Med
1999
, vol. 
5
 
5
(pg. 
512
-
517
)
4
Siliciano
 
JD
Kajdas
 
J
Finzi
 
D
, et al. 
Long-term follow-up studies confirm the stability of the latent reservoir for HIV-1 in resting CD4+ T cells.
Nat Med
2003
, vol. 
9
 
6
(pg. 
727
-
728
)
5
US Department of Health and Human Services. HIV/AIDS care continuum. https://www.aids.gov/federal-resources/policies/care-continuum/. Accessed January 12, 2016
6
Kelley
 
CF
Kitchen
 
CM
Hunt
 
PW
, et al. 
Incomplete peripheral CD4+ cell count restoration in HIV-infected patients receiving long-term antiretroviral treatment.
Clin Infect Dis
2009
, vol. 
48
 
6
(pg. 
787
-
794
)
7
Burnett
 
JC
Zaia
 
JA
Rossi
 
JJ
Creating genetic resistance to HIV.
Curr Opin Immunol
2012
, vol. 
24
 
5
(pg. 
625
-
632
)
8
Lam
 
S
Bollard
 
C
T-cell therapies for HIV.
Immunotherapy
2013
, vol. 
5
 
4
(pg. 
407
-
414
)
9
Mylvaganam
 
GH
Silvestri
 
G
Amara
 
RR
HIV therapeutic vaccines: moving towards a functional cure.
Curr Opin Immunol
2015
, vol. 
35
 (pg. 
1
-
8
)
10
Morgan
 
RA
Walker
 
R
Carter
 
CS
, et al. 
Preferential survival of CD4+ T lymphocytes engineered with anti-human immunodeficiency virus (HIV) genes in HIV-infected individuals.
Hum Gene Ther
2005
, vol. 
16
 
9
(pg. 
1065
-
1074
)
11
Li
 
MJ
Kim
 
J
Li
 
S
, et al. 
Long-term inhibition of HIV-1 infection in primary hematopoietic cells by lentiviral vector delivery of a triple combination of anti-HIV shRNA, anti-CCR5 ribozyme, and a nucleolar-localizing TAR decoy.
Mol Ther
2005
, vol. 
12
 
5
(pg. 
900
-
909
)
12
Podsakoff
 
GM
Engel
 
BC
Carbonaro
 
DA
, et al. 
Selective survival of peripheral blood lymphocytes in children with HIV-1 following delivery of an anti-HIV gene to bone marrow CD34(+) cells.
Mol Ther
2005
, vol. 
12
 
1
(pg. 
77
-
86
)
13
Levine
 
BL
Humeau
 
LM
Boyer
 
J
, et al. 
Gene transfer in humans using a conditionally replicating lentiviral vector.
Proc Natl Acad Sci USA
2006
, vol. 
103
 
46
(pg. 
17372
-
17377
)
14
van Lunzen
 
J
Glaunsinger
 
T
Stahmer
 
I
, et al. 
Transfer of autologous gene-modified T cells in HIV-infected patients with advanced immunodeficiency and drug-resistant virus.
Mol Ther
2007
, vol. 
15
 
5
(pg. 
1024
-
1033
)
15
Mitsuyasu
 
RT
Merigan
 
TC
Carr
 
A
, et al. 
Phase 2 gene therapy trial of an anti-HIV ribozyme in autologous CD34+ cells.
Nat Med
2009
, vol. 
15
 
3
(pg. 
285
-
292
)
16
DiGiusto
 
DL
Krishnan
 
A
Li
 
L
, et al. 
RNA-based gene therapy for HIV with lentiviral vector-modified CD34(+) cells in patients undergoing transplantation for AIDS-related lymphoma.
Sci Transl Med
2010
, vol. 
2
 
36
pg. 
36ra43
 
17
Nienhuis
 
AW
Dunbar
 
CE
Sorrentino
 
BP
Genotoxicity of retroviral integration in hematopoietic cells.
Mol Ther
2006
, vol. 
13
 
6
(pg. 
1031
-
1049
)
18
Trobridge
 
GD
Genotoxicity of retroviral hematopoietic stem cell gene therapy.
Expert Opin Biol Ther
2011
, vol. 
11
 
5
(pg. 
581
-
593
)
19
Belfort
 
M
Roberts
 
RJ
Homing endonucleases: keeping the house in order.
Nucleic Acids Res
1997
, vol. 
25
 
17
(pg. 
3379
-
3388
)
20
Gaj
 
T
Gersbach
 
CA
Barbas
 
CF
ZFN, TALEN, and CRISPR/Cas-based methods for genome engineering.
Trends Biotechnol
2013
, vol. 
31
 
7
(pg. 
397
-
405
)
21
Saleh-Gohari
 
N
Helleday
 
T
Conservative homologous recombination preferentially repairs DNA double-strand breaks in the S phase of the cell cycle in human cells.
Nucleic Acids Res
2004
, vol. 
32
 
12
(pg. 
3683
-
3688
)
22
Karanam
 
K
Kafri
 
R
Loewer
 
A
Lahav
 
G
Quantitative live cell imaging reveals a gradual shift between DNA repair mechanisms and a maximal use of HR in mid S phase.
Mol Cell
2012
, vol. 
47
 
2
(pg. 
320
-
329
)
23
Hoffmann
 
C
The epidemiology of HIV coreceptor tropism.
Eur J Med Res
2007
, vol. 
12
 
9
(pg. 
385
-
390
)
24
Samson
 
M
Libert
 
F
Doranz
 
BJ
, et al. 
Resistance to HIV-1 infection in caucasian individuals bearing mutant alleles of the CCR-5 chemokine receptor gene.
Nature
1996
, vol. 
382
 
6593
(pg. 
722
-
725
)
25
Huang
 
Y
Paxton
 
WA
Wolinsky
 
SM
, et al. 
The role of a mutant CCR5 allele in HIV-1 transmission and disease progression.
Nat Med
1996
, vol. 
2
 
11
(pg. 
1240
-
1243
)
26
Liu
 
R
Paxton
 
WA
Choe
 
S
, et al. 
Homozygous defect in HIV-1 coreceptor accounts for resistance of some multiply-exposed individuals to HIV-1 infection.
Cell
1996
, vol. 
86
 
3
(pg. 
367
-
377
)
27
Hütter
 
G
Nowak
 
D
Mossner
 
M
, et al. 
Long-term control of HIV by CCR5 Delta32/Delta32 stem-cell transplantation.
N Engl J Med
2009
, vol. 
360
 
7
(pg. 
692
-
698
)
28
Allers
 
K
Hütter
 
G
Hofmann
 
J
, et al. 
Evidence for the cure of HIV infection by CCR5Δ32/Δ32 stem cell transplantation.
Blood
2011
, vol. 
117
 
10
(pg. 
2791
-
2799
)
29
Henrich
 
TJ
Hu
 
Z
Li
 
JZ
, et al. 
Long-term reduction in peripheral blood HIV type 1 reservoirs following reduced-intensity conditioning allogeneic stem cell transplantation.
J Infect Dis
2013
, vol. 
207
 
11
(pg. 
1694
-
1702
)
30
Cannon
 
PM
Kohn
 
DB
Kiem
 
HP
HIV eradication--from Berlin to Boston.
Nat Biotechnol
2014
, vol. 
32
 
4
(pg. 
315
-
316
)
31
Cannon
 
P
June
 
C
Chemokine receptor 5 knockout strategies.
Curr Opin HIV AIDS
2011
, vol. 
6
 
1
(pg. 
74
-
79
)
32
Yang
 
AG
Bai
 
X
Huang
 
XF
Yao
 
C
Chen
 
S
Phenotypic knockout of HIV type 1 chemokine coreceptor CCR-5 by intrakines as potential therapeutic approach for HIV-1 infection.
Proc Natl Acad Sci USA
1997
, vol. 
94
 
21
(pg. 
11567
-
11572
)
33
Bai
 
J
Gorantla
 
S
Banda
 
N
Cagnon
 
L
Rossi
 
J
Akkina
 
R
Characterization of anti-CCR5 ribozyme-transduced CD34+ hematopoietic progenitor cells in vitro and in a SCID-hu mouse model in vivo.
Mol Ther
2000
, vol. 
1
 
3
(pg. 
244
-
254
)
34
Cordelier
 
P
Kulkowsky
 
JW
Ko
 
C
, et al. 
Protecting from R5-tropic HIV: individual and combined effectiveness of a hammerhead ribozyme and a single-chain Fv antibody that targets CCR5.
Gene Ther
2004
, vol. 
11
 
22
(pg. 
1627
-
1637
)
35
Qin
 
XF
An
 
DS
Chen
 
IS
Baltimore
 
D
Inhibiting HIV-1 infection in human T cells by lentiviral-mediated delivery of small interfering RNA against CCR5.
Proc Natl Acad Sci USA
2003
, vol. 
100
 
1
(pg. 
183
-
188
)
36
Anderson
 
J
Akkina
 
R
HIV-1 resistance conferred by siRNA cosuppression of CXCR4 and CCR5 coreceptors by a bispecific lentiviral vector.
AIDS Res Ther
2005
, vol. 
2
 
1
pg. 
1
 
37
Corrigan-Curay
 
J
O’Reilly
 
M
Kohn
 
DB
, et al. 
Genome editing technologies: defining a path to clinic.
Mol Ther
2015
, vol. 
23
 
5
(pg. 
796
-
806
)
38
Perez
 
EE
Wang
 
J
Miller
 
JC
, et al. 
Establishment of HIV-1 resistance in CD4+ T cells by genome editing using zinc-finger nucleases.
Nat Biotechnol
2008
, vol. 
26
 
7
(pg. 
808
-
816
)
39
Holt
 
N
Wang
 
J
Kim
 
K
, et al. 
Human hematopoietic stem/progenitor cells modified by zinc-finger nucleases targeted to CCR5 control HIV-1 in vivo.
Nat Biotechnol
2010
, vol. 
28
 
8
(pg. 
839
-
847
)
40
Maier
 
DA
Brennan
 
AL
Jiang
 
S
, et al. 
Efficient clinical scale gene modification via zinc finger nuclease-targeted disruption of the HIV co-receptor CCR5.
Hum Gene Ther
2013
, vol. 
24
 
3
(pg. 
245
-
258
)
41
Tebas
 
P
Stein
 
D
Tang
 
WW
, et al. 
Gene editing of CCR5 in autologous CD4 T cells of persons infected with HIV.
N Engl J Med
2014
, vol. 
370
 
10
(pg. 
901
-
910
)
42
US National Institutes of Health. Autologous T-cells genetically modified at the CCR5 gene by zinc finger nucleases SB-728 for HIV (zinc-finger). https://clinicaltrials.gov/ct2/show/NCT00842634. Accessed January 12, 2016
43
Ando
 
D
Lalezari
 
J
Blick
 
G
, et al. 
 
HIV protected autologous zinc finger nuclease CCR5 modified CD4 cells (SB-728-T) reduce Viral Load (VL) in HIV subjects during treatment interruption (TI): correlates of effect, and effect of cytoxan conditioning. ICAAC 2014. http://www.natap.org/2014/ICAAC/ICAAC_36.htm. Accessed January 12, 2016
44
Ando
 
D
 
Functional control of viremia in CCR5-D31 heterozygous HIV+ subjects following adoptive transfer of zinc finger nuclease CCR5 modified autulogous CD4 T-cells. ESGCT 2013. http://www.natap.org/2013/HIV/103013_01.htm. Accessed January 12, 2016
45
Tsai
 
SQ
Zheng
 
Z
Nguyen
 
NT
, et al. 
GUIDE-seq enables genome-wide profiling of off-target cleavage by CRISPR-Cas nucleases.
Nat Biotechnol
2015
, vol. 
33
 
2
(pg. 
187
-
197
)
46
Frock
 
RL
Hu
 
J
Meyers
 
RM
Ho
 
YJ
Kii
 
E
Alt
 
FW
Genome-wide detection of DNA double-stranded breaks induced by engineered nucleases.
Nat Biotechnol
2015
, vol. 
33
 
2
(pg. 
179
-
186
)
47
Smith
 
MW
Carrington
 
M
Winkler
 
C
, et al. 
CCR2 chemokine receptor and AIDS progression.
Nat Med
1997
, vol. 
3
 
10
(pg. 
1052
-
1053
)
48
Li
 
L
Krymskaya
 
L
Wang
 
J
, et al. 
Genomic editing of the HIV-1 coreceptor CCR5 in adult hematopoietic stem and progenitor cells using zinc finger nucleases.
Mol Ther
2013
, vol. 
21
 
6
(pg. 
1259
-
1269
)
49
Wang
 
J
Exline
 
CM
DeClercq
 
JJ
, et al. 
Homology-driven genome editing in hematopoietic stem and progenitor cells using ZFN mRNA and AAV6 donors.
Nat Biotechnol
2015
, vol. 
33
 
12
(pg. 
1256
-
1263
)
50
Hofer
 
U
Henley
 
JE
Exline
 
CM
Mulhern
 
O
Lopez
 
E
Cannon
 
PM
Pre-clinical modeling of CCR5 knockout in human hematopoietic stem cells by zinc finger nucleases using humanized mice.
J Infect Dis
2013
, vol. 
208
 
Suppl 2
(pg. 
S160
-
S164
)
51
Holt
 
N
Exline
 
CM
Mulhern
 
O
, et al. 
Zinc finger nuclease editing of hematopoietic stem cells as an anti-HIV therapy.
 
In: Poluektova LY, Garcia-Martinez JV, Koyanagi Y, Manz MG, Tager AM, eds. Humanized Mice for HIV Research. New York: Springer; 2014:407-416
52
Cannon
 
PM
Henley
 
J
Wood
 
T
, et al. 
Electroporation of ZFN mRNA enables efficient CCR5 gene disruption in mobilized blood hematopoietic stem cells at clinical scale.
Mol Ther
2013
, vol. 
21
 
S1
(pg. 
S71
-
S72
)
53
Genovese
 
P
Schiroli
 
G
Escobar
 
G
, et al. 
Targeted genome editing in human repopulating haematopoietic stem cells.
Nature
2014
, vol. 
510
 
7504
(pg. 
235
-
240
)
54
Choi
 
KD
Yu
 
J
Smuga-Otto
 
K
, et al. 
Hematopoietic and endothelial differentiation of human induced pluripotent stem cells.
Stem Cells
2009
, vol. 
27
 
3
(pg. 
559
-
567
)
55
Yao
 
Y
Nashun
 
B
Zhou
 
T
, et al. 
Generation of CD34+ cells from CCR5-disrupted human embryonic and induced pluripotent stem cells.
Hum Gene Ther
2012
, vol. 
23
 
2
(pg. 
238
-
242
)
56
Ye
 
L
Wang
 
J
Beyer
 
AI
, et al. 
Seamless modification of wild-type induced pluripotent stem cells to the natural CCR5Δ32 mutation confers resistance to HIV infection.
Proc Natl Acad Sci USA
2014
, vol. 
111
 
26
(pg. 
9591
-
9596
)
57
Mussolino
 
C
Alzubi
 
J
Fine
 
EJ
, et al. 
TALENs facilitate targeted genome editing in human cells with high specificity and low cytotoxicity.
Nucleic Acids Res
2014
, vol. 
42
 
10
(pg. 
6762
-
6773
)
58
Sather
 
BD
Romano Ibarra
 
GS
Sommer
 
K
, et al. 
Efficient modification of CCR5 in primary human hematopoietic cells using a megaTAL nuclease and AAV donor template.
Sci Transl Med
2015
, vol. 
7
 
307
pg. 
307ra156
 
59
Mandal
 
PK
Ferreira
 
LM
Collins
 
R
, et al. 
Efficient ablation of genes in human hematopoietic stem and effector cells using CRISPR/Cas9.
Cell Stem Cell
2014
, vol. 
15
 
5
(pg. 
643
-
652
)
60
Li
 
C
Guan
 
X
Du
 
T
, et al. 
Inhibition of HIV-1 infection of primary CD4+ T-cells by gene editing of CCR5 using adenovirus-delivered CRISPR/Cas9.
J Gen Virol
2015
, vol. 
96
 
8
(pg. 
2381
-
2393
)
61
Hendel
 
A
Bak
 
RO
Clark
 
JT
, et al. 
Chemically modified guide RNAs enhance CRISPR-Cas genome editing in human primary cells.
Nat Biotechnol
2015
, vol. 
33
 
9
(pg. 
985
-
989
)
62
Connor
 
RI
Sheridan
 
KE
Ceradini
 
D
Choe
 
S
Landau
 
NR
Change in coreceptor use correlates with disease progression in HIV-1--infected individuals.
J Exp Med
1997
, vol. 
185
 
4
(pg. 
621
-
628
)
63
Peled
 
A
Petit
 
I
Kollet
 
O
, et al. 
Dependence of human stem cell engraftment and repopulation of NOD/SCID mice on CXCR4.
Science
1999
, vol. 
283
 
5403
(pg. 
845
-
848
)
64
Wilen
 
CB
Wang
 
J
Tilton
 
JC
, et al. 
Engineering HIV-resistant human CD4+ T cells with CXCR4-specific zinc-finger nucleases.
PLoS Pathog
2011
, vol. 
7
 
4
pg. 
e1002020
 
65
Yuan
 
J
Wang
 
J
Crain
 
K
, et al. 
Zinc-finger nuclease editing of human cxcr4 promotes HIV-1 CD4(+) T cell resistance and enrichment.
Mol Ther
2012
, vol. 
20
 
4
(pg. 
849
-
859
)
66
Didigu
 
CA
Wilen
 
CB
Wang
 
J
, et al. 
Simultaneous zinc-finger nuclease editing of the HIV coreceptors ccr5 and cxcr4 protects CD4+ T cells from HIV-1 infection.
Blood
2014
, vol. 
123
 
1
(pg. 
61
-
69
)
67
Voit
 
RA
McMahon
 
MA
Sawyer
 
SL
Porteus
 
MH
Generation of an HIV resistant T-cell line by targeted “stacking” of restriction factors.
Mol Ther
2013
, vol. 
21
 
4
(pg. 
786
-
795
)
68
Malim
 
MH
Bieniasz
 
PD
HIV restriction factors and mechanisms of evasion.
Cold Spring Harb Perspect Med
2012
, vol. 
2
 
5
pg. 
a006940
 
69
Xu
 
H
Svarovskaia
 
ES
Barr
 
R
, et al. 
A single amino acid substitution in human APOBEC3G antiretroviral enzyme confers resistance to HIV-1 virion infectivity factor-induced depletion.
Proc Natl Acad Sci USA
2004
, vol. 
101
 
15
(pg. 
5652
-
5657
)
70
Yap
 
MW
Nisole
 
S
Stoye
 
JP
A single amino acid change in the SPRY domain of human Trim5alpha leads to HIV-1 restriction.
Curr Biol
2005
, vol. 
15
 
1
(pg. 
73
-
78
)
71
Gupta
 
RK
Hué
 
S
Schaller
 
T
Verschoor
 
E
Pillay
 
D
Towers
 
GJ
Mutation of a single residue renders human tetherin resistant to HIV-1 Vpu-mediated depletion.
PLoS Pathog
2009
, vol. 
5
 
5
pg. 
e1000443
 
72
Chinn
 
LW
Tang
 
M
Kessing
 
BD
, et al. 
Genetic associations of variants in genes encoding HIV-dependency factors required for HIV-1 infection.
J Infect Dis
2010
, vol. 
202
 
12
(pg. 
1836
-
1845
)
73
Murali
 
TM
Dyer
 
MD
Badger
 
D
Tyler
 
BM
Katze
 
MG
Network-based prediction and analysis of HIV dependency factors.
PLOS Comput Biol
2011
, vol. 
7
 
9
pg. 
e1002164
 
74
Cleret-Buhot
 
A
Zhang
 
Y
Planas
 
D
, et al. 
Identification of novel HIV-1 dependency factors in primary CCR4(+)CCR6(+)Th17 cells via a genome-wide transcriptional approach.
Retrovirology
2015
, vol. 
12
 
1
pg. 
102
 
75
Fellay
 
J
Shianna
 
KV
Ge
 
D
, et al. 
A whole-genome association study of major determinants for host control of HIV-1.
Science
2007
, vol. 
317
 
5840
(pg. 
944
-
947
)
76
Lombardo
 
A
Genovese
 
P
Beausejour
 
CM
, et al. 
Gene editing in human stem cells using zinc finger nucleases and integrase-defective lentiviral vector delivery.
Nat Biotechnol
2007
, vol. 
25
 
11
(pg. 
1298
-
1306
)
77
Hendel
 
A
Kildebeck
 
EJ
Fine
 
EJ
, et al. 
Quantifying genome-editing outcomes at endogenous loci with SMRT sequencing.
Cell Reports
2014
, vol. 
7
 
1
(pg. 
293
-
305
)
78
Urnov
 
FD
Miller
 
JC
Lee
 
YL
, et al. 
Highly efficient endogenous human gene correction using designed zinc-finger nucleases.
Nature
2005
, vol. 
435
 
7042
(pg. 
646
-
651
)
79
Zielske
 
SP
Reese
 
JS
Lingas
 
KT
Donze
 
JR
Gerson
 
SL
In vivo selection of MGMT(P140K) lentivirus-transduced human NOD/SCID repopulating cells without pretransplant irradiation conditioning.
J Clin Invest
2003
, vol. 
112
 
10
(pg. 
1561
-
1570
)
80
Davis
 
BM
Humeau
 
L
Dropulic
 
B
In vivo selection for human and murine hematopoietic cells transduced with a therapeutic MGMT lentiviral vector that inhibits HIV replication.
Mol Ther
2004
, vol. 
9
 
2
(pg. 
160
-
172
)
81
Sarkar
 
I
Hauber
 
I
Hauber
 
J
Buchholz
 
F
HIV-1 proviral DNA excision using an evolved recombinase.
Science
2007
, vol. 
316
 
5833
(pg. 
1912
-
1915
)
82
Karpinski
 
J
Chemnitz
 
J
Hauber
 
I
, et al. 
Universal Tre (uTre) recombinase specifically targets the majority of HIV-1 isolates.
J Int AIDS Soc
2014
, vol. 
17
 
4 Suppl 3
pg. 
19706
 
83
Qu
 
X
Wang
 
P
Ding
 
D
, et al. 
Zinc-finger-nucleases mediate specific and efficient excision of HIV-1 proviral DNA from infected and latently infected human T cells.
Nucleic Acids Res
2013
, vol. 
41
 
16
(pg. 
7771
-
7782
)
84
Ebina
 
H
Misawa
 
N
Kanemura
 
Y
Koyanagi
 
Y
Harnessing the CRISPR/Cas9 system to disrupt latent HIV-1 provirus.
Sci Rep
2013
, vol. 
3
 pg. 
2510
 
85
Hu
 
W
Kaminski
 
R
Yang
 
F
, et al. 
RNA-directed gene editing specifically eradicates latent and prevents new HIV-1 infection.
Proc Natl Acad Sci USA
2014
, vol. 
111
 
31
(pg. 
11461
-
11466
)
86
Zhu
 
W
Lei
 
R
Le Duff
 
Y
, et al. 
The CRISPR/Cas9 system inactivates latent HIV-1 proviral DNA.
Retrovirology
2015
, vol. 
12
 pg. 
22
 
87
Ebina
 
H
Kanemura
 
Y
Misawa
 
N
, et al. 
A high excision potential of TALENs for integrated DNA of HIV-based lentiviral vector.
PLoS One
2015
, vol. 
10
 
3
pg. 
e0120047
 
88
Aubert
 
M
Ryu
 
BY
Banks
 
L
Rawlings
 
DJ
Scharenberg
 
AM
Jerome
 
KR
Successful targeting and disruption of an integrated reporter lentivirus using the engineered homing endonuclease Y2 I-AniI.
PLoS One
2011
, vol. 
6
 
2
pg. 
e16825
 
89
Münch
 
RC
Muth
 
A
Muik
 
A
, et al. 
Off-target-free gene delivery by affinity-purified receptor-targeted viral vectors.
Nat Commun
2015
, vol. 
6
 pg. 
6246
 
90
Zhou
 
Q
Uhlig
 
KM
Muth
 
A
, et al. 
Exclusive transduction of human CD4+ T cells upon systemic delivery of CD4-targeted lentiviral vectors.
J Immunol
2015
, vol. 
195
 
5
(pg. 
2493
-
2501
)
Sign in via your Institution