AAV Vector Immunogenicity in Humans: A Long Journey to Successful Gene Transfer (2024)

As a library, NLM provides access to scientific literature. Inclusion in an NLM database does not imply endorsem*nt of, or agreement with, the contents by NLM or the National Institutes of Health.
Learn more: PMC Disclaimer | PMC Copyright Notice

AAV Vector Immunogenicity in Humans: A Long Journey to Successful Gene Transfer (1)

Molecular Therapy

Mol Ther. 2020 Mar 4; 28(3): 723–746.

Published online 2020 Jan 10. doi:10.1016/j.ymthe.2019.12.010

PMCID: PMC7054726

PMID: 31972133

Author information Copyright and License information PMC Disclaimer

Abstract

Gene therapy with adeno-associated virus (AAV) vectors has demonstrated safety and long-term efficacy in a number of trials across target organs, including eye, liver, skeletal muscle, and the central nervous system. Since the initial evidence that AAV vectors can elicit capsid Tcell responses in humans, which can affect the duration of transgene expression, much progress has been made in understanding and modulating AAV vector immunogenicity. It is now well established that exposure to wild-type AAV results in priming of the immune system against the virus, with development of both humoral and Tcell immunity. Aside from the neutralizing effect of antibodies, the impact of pre-existing immunity to AAV on gene transfer is still poorly understood. Herein, we review data emerging from clinical trials across a broad range of gene therapy applications. Common features of immune responses to AAV can be found, suggesting, for example, that vector immunogenicity is dose-dependent, and that innate immunity plays an important role in the outcome of gene transfer. A range of host-specific factors are also likely to be important, and a comprehensive understanding of the mechanisms driving AAV vector immunogenicity in humans will be key to unlocking the full potential of invivo gene therapy.

Keywords: AAV vectors, immune responses, T cells, antibody responses, gene therapy

Graphical Abstract

AAV vectors are broadly utilized tools for invivo gene therapy. Yet, the human immune system poses important challenges to the use of this platform. The complex interplay of innate and adaptive immunity, to both vector components and transgene product, appears to be a key determinant of the outcomes in gene transfer trials.

Introduction

Recombinant adeno-associated virus (AAV) vectors are derived from small, non-enveloped, 4.7-kb DNA dependoviruses belonging to the Parvoviridae family. During the past several years, invivo gene therapy with AAV vectors has demonstrated the potential of correcting genetic disorders in a permanent manner by delivering a functional copy of a gene into the nucleus of somatic cells in affected tissues. The transferred gene, or transgene, compensates for genetic mutations underlying inherited genetic disorders. AAV vectors are particularly attractive as invivo gene delivery tools, as they are mostly non-integrative and can transduce a wide variety of terminally differentiated tissues, driving long-term transgene expression,1, 2, 3, 4, 5 while they are inefficient at transducing antigen-presenting cells (APCs) and have a low immunogenicity profile.6,7 Despite this, immune responses encountered in humans undergoing gene transfer with AAV vectors have been an important obstacle to the advancement of the field (Figure1).

AAV Vector Immunogenicity in Humans: A Long Journey to Successful Gene Transfer (3)

Host Immune Responses against AAV Vectors

Prior to vector administration, humans are exposed to wild-type AAV and therefore can develop both humoral and Tcell-mediated immunity to the vector. Exposure to wild-type AAV can occur years prior to gene transfer, and together with host-specific factors can determine the overall immunological context of AAV vector delivery. Immediately after vector delivery, the vector in its components can trigger innate immune recognition. While no evidence of severe systemic inflammation has been observed in AAV trials immediately after vector delivery, some episodes of pyrexia have been documented, as well as toxicities potentially associated with complement activation. Later after vector administration, anti-capsid antibodies are produced and persist for several years after gene transfer. Capsid Tcell activation has also been documented in several trials, in some cases correlating directly with loss of transgene expression. Transgene immune responses are also a potential immune-related risk in gene therapy, although to date they have been documented only in isolated trials.

The wild-type (WT) AAV is highly prevalent in the human population,8 although exposure to this virus has not been clearly associated with any clinical pathology or disease.9 After primary infection, WT AAV genomes can persist for years in host cells, either episomally orintegrated within the host DNA, and be reactivated by a helper virus, such as adenovirus, herpesvirus, human papillomavirus, andvaccinia virus,10, 11, 12, 13 or a genotoxic reagent. Recent studies have linked the integration of the WT AAV genome to the development of hepatocellular carcinoma,14,15 although to date no evidence of genotoxicity has emerged from the long-term follow-up of subjects enrolled in gene transfer studies.

Several different natural AAV serotypes have been isolated in nature,16 which differ in the sequence of their capsid. The capsid serotype and the presence of a specific receptor on the host cellsdetermine the tropism of each AAV serotype for a tissue (Table1),16,17 a property that makes AAVs versatile vectors adaptable to a broad range of therapeutic applications. AAV vectors can be manufactured according to various methods,18 although the most common is bytransfection of the HEK293 cell line with three different DNA plasmids encoding the vector genome, the rep and cap genes derived from a specific AAV serotype, and a helper plasmid.19

Table 1

Receptors and Preferential Tissue Tropism of Natural AAV Vectors

SerotypeSourceGlycan ReceptorCo-receptor/OtherExamples of Tissue Tropism
AAV1non-human primateN-linked sialic acidunknownskeletal muscle, lung, CNS, retina, pancreas
AAV2humanHSPGFGFR1, HGFR, LamR, CD9 tetraspaninsmooth muscle, skeletal muscle, CNS, liver, kidney
AAV3non-human primateHSPGFGFR1, HGFR, LamRhepatocarcinoma, skeletal muscle, inner ear
AAV4non-human primateO-linked sialic acidunknownCNS, retina
AAV5humanN-linked sialic acidPDGFRskeletal muscle, CNS, lung, retina, liver
AAV6humanN-linked sialic acid, HSPGEGFRskeletal muscle, heart, lung, bone marrow
AAV7non-human primateunknownunknownskeletal muscle, retina, CNS
AAV8non-human primateunknownLamRliver, skeletal muscle, CNS, retina, pancreas, heart
AAV9non-human primateN-linked galactoseLamRliver, heart, brain, skeletal muscle, lungs, pancreas, kidney
AAV10non-human primateunknownunknownliver

Pre-existing Immunity to WT AAV in Humans

Pre-existing Humoral Immunity

Several studies have investigated the seroprevalence of neutralizing antibodies directed against WT AAV in humans.20, 21, 22, 23, 24, 25 Seroprevalence varies geographically, with anti-AAV2 neutralizing antibodies displaying the highest prevalence, ranging from 30% to 60% of the population. Due to the broad cross-reactivity between AAV serotypes,26 neutralizing antibodies recognizing virtually all serotypes can be found in almost all subjects.27 This cross-reactivity reflects the amino acid sequence and structural hom*ology across capsids of different AAV serotypes.28

The prevalence of total anti-AAV antibodies is close to 70% of thepopulation for AAV1 and AAV2, 45% for AAV6 and AAV9, and 38% for AAV8.20 Importantly, titers of anti-AAV immunoglobulin G (IgG) antibodies correlate significantly with titers of anti-AAVneutralizing antibodies,27,29 although some individuals carry non-neutralizing anti-AAV IgG.30 Aside from the neutralizing effect on AAV vectors,31, 32, 33 relatively little is known about the effect of antibodies on the vector tropism and immunogenicity. Systemic delivery of AAV vectors in the presence of neutralizing antibodies has been shown to result in the accumulation of vector genomes in lymphoid organs.30,34 Conversely, in the presence of binding, non-neutralizing antibodies, the transduction efficiency of organs such as the liver appears to be enhanced.30 An additional role of antibodies as mediators of toxicities associated with complement activation35 in gene transfer trials is being investigated (videinfra), although it is known that AAV vectors interact with complement proteins.36,37

IgG1 appears to be the predominant immunoglobulin subclass in WTAAV seropositive individuals,20,38 although some subjects carry high levels of IgG2 and IgG3. Titers of IgG1, IgG2, and IgM are well correlated with the level of neutralizing antibodies (Nabs), which is not the case for IgG3 and IgG4.38 Similarly, in subjects undergoing AAV vector gene transfer, the development of high-titer IgG1 antibodies together with NAbs has been documented.38 In contrast, levels of IgG3 have been found to be correlated with the detection of Tcell reactivity to AAV vectors.39

Pre-existing Cellular Immunity

Since the initial findings of Tcell-mediated immunity in AAV gene transfer trials,32,40,41 the scientific community has focused on pre-existing anti-AAV cellular immunity and its potential impact on AAV-based gene transfer.42, 43, 44

The prevalence of Tcells directed against AAV1 and AAV2 in the general population has been investigated through a variety of functional assays, which include enzyme-linked immunospot (ELISPOT)45, 46, 47 and flow cytometry-based assays.42,45,47 Although the prevalence can vary across studies, depending on the sensitivity of the assay, capsid-specific cellular responses are less frequent, or less detectable, than humoral responses. Several studies40,45,47 have pointed out the lack of correlation between detection of Tcells secreting interferon (IFN)γ in response to capsid antigen and thepresence of anti-AAV antibodies in serum. More recently, we identified a correlation between detection of circulating capsid-specific memory CD8+ Tcells secreting tumor necrosis factor (TNF)-α and the presence of anti-AAV antibodies,42 possibly indicating that IFNγ is not the main signature cytokine of Tcell-mediated immune responses to AAV.

Capsid-specific Tcells are highly cross-reactive45 and can be detected more frequently in splenocytes compared to peripheral blood mononuclear cells (PBMCs), suggesting that AAV-specific memory Tcells might fail to recirculate in peripheral blood and preferentially home to lymphoid organs.40,45 In addition, a higher prevalence of Tcell responses in PBMCs or splenocytes is observed after several rounds of invitro expansion, suggesting that the frequency of AAV-specific Tcells is too low to be efficiently detected exvivo.40,45 The fact that flow cytometry-based assessment of differentiation markers evidenced that most AAV-specific Tcells exhibit a memory phenotype40,42,43,47 suggests that they might arise during infancy afternaturally occurring WT AAV infections, and persist throughout lifetime as a pool of memory Tcells in secondary lymphoid organs.This hypothesis is consistent with the detection of reactive Tcells in splenocytes from adults and, at a lower frequency, from children.40,45

AAV-specific memory Tcells have been shown to produce IFNγ, interleukin (IL)-2, and TNF-α and to present a cytotoxic phenotype characterized by the expression of granzyme B and CD107a degranulation markers.42,43,45,47 We recently identified two patterns of cellular responses to AAV depending on the serology of patients. By conventional and CyTOF mass cytometry,48 we observed that exposure of human PBMCs to AAV capsid epitopes induced the activation of CD8+ Tcells with an effector memory phenotype, whichsecreted predominantly TNF-α, but also cytolytic granules containing granzyme B and CD107a.42 Alternatively, in seronegative patients we observed transient activation of natural killer (NK) cells,but not naive CD8+ Tcells. These NK cells secreted both IFNγ and TNF-α, but they did not show a cytotoxic phenotype. The role of NK cells as effectors of cellular immune responses to AAV remains to be characterized.

Unlike for humoral immunity, the role of pre-existing Tcell responses directed against capsid epitopes in the outcome of gene transfer is not entirely understood. The timing of detection of Tcell reactivity to AAV in gene transfer trials has been inconsistent withthat of a memory recall response. However, the current knowledge on immune responses in the context of an ongoing viral infection may not entirely apply to the setting of gene therapy, in which, for example, large quantities or non-replicating virions are infused directly into the bloodstream.

Induction of Immune Responses against AAV Vectors

Innate Capsid Immunogenicity

Several complex processes contribute to generating an immune response against an offending pathogen. Because AAV vectors lack any coding viral sequence, the main sources of foreign antigens brought in during gene transfer, aside from contaminant carryovers from the production and purification process, are derived from initial input of viral capsid and from the transgene product. The DNA component of AAV vectors, and possibly double-stranded RNA (dsRNA) produced by the vector itself,49 can also act as an adjuvant, concurring in the activation of innate immunity along with other host-specific factors (Figure2).

AAV Vector Immunogenicity in Humans: A Long Journey to Successful Gene Transfer (4)

Factors that Influence AAV Vector Immunogenicity

The capsid, its genome, and the transgene product are the main potential immunogenic components of AAV vectors. Production of dsRNA driven by the promoter activity of ITRs can also act as a trigger for innate immunity. Additional host-dependent and vector-dependent factors can modulate the overall vector immunogenicity. These factors are mostly poorly understood, although the presence of innate immunity activators such as CpG and vector dose seem to correlate with vector-related immunotoxicities in some trials.

In recent years, innate immunity to AAV has gained increasing attention as one of the possible triggers of immune-mediated toxicities observed in clinical trials, although the overall lack of clinical evidence (e.g., detection of proinflammatory cytokines in the circulation) has been a challenge to clearly establish a direct causative relationship.

Innate immunity mounts rapidly, is non-specific, and does not result in immunological memory. Innate immune responses are initiated through the recognition of pathogen-associated molecular patterns (PAMPs), exhibited on pathogens, by PRRs (pattern recognition receptors) expressed at the surface or within immune cells. These PRRs recognize viral nucleic acids, as well as membrane glycoproteins, or even chemical messengers. Through a variety of signaling pathways, the engagement of PRRs mainly leads to the activation ofNF-κB (nuclear factor κB) and IRF (IFN-regulatory factor) transcription factors, both of which play a central role in inducing the expression of pro-inflammatory cytokines or type I IFNs, respectively.50 Type I IFNs have been described as being important forthe induction of anti-capsid CD8+ Tcell responses.51 In the preclinical setting, blocking the pathways of activation of innate immune responses has been shown to prevent anti-capsid cytotoxic51,52 and humoral responses42invivo. In non-parenchymal liver cells, including Kupffer cells and liver sinusoidal endothelial cells (LSECs), the viral capsid has been seen to activate innate immunity mainly through binding to Toll-like receptor (TLR)2 expressed on the cell surface.53 Moreover, the double-stranded DNA vector genome, and in particular unmethylated CpG motifs, are recognized by the endosomal TLR9 in Kupffer cells,54 peripheral plasmacytoid dendritic cells(pDCs),52,55 and monocyte-derived DCs.56 TLR9 engagement has been associated with enhanced capsid antigen presentation into major histocompatibility complex (MHC) class I and subsequent capsid-specific CD8+ Tcell activation.44,52,57 Also, the MyD88 (myeloid differentiating factor 88)-TLR9 pathway has been shown to mediate the induction of immune responses to liver- and muscle-targeted transgenes.56,58

In addition to the vector DNA, a recent study suggested a possible contribution of dsRNA to the induction of innate immunity to AAV.49 This would explain why cellular responses are sometimes initiated weeks after vector administration in clinical trials.5,41 According to this study, the promoter activity of the inverted terminalrepeats (ITRs) flanking the transgene expression cassette could potentially drive the production of dsRNA, which in turn wouldstimulate the MDA5 sensor in human hepatocytes transduced with AAV, leading to the expression of type I IFN. Interestingly, blockade of MDA5 decreased the IFN response and improved transgene expression in transduced cells invitro.49

Adaptive Immunity

Adaptive immunity occurs after innate immunity and allows theantigen-specific recognition and elimination of pathogens, followed by the establishment of immunological memory. During the establishment of an adaptive response, T and B lymphocytes become activated after recognizing an antigen presented by APCs.59 After activation, lymphocytes expand and differentiate into effector cells and mediate the elimination of antigens through the induction of humoral or cytotoxic responses. After clearing the antigen, the adaptive immune response is followed by a contraction phase and the generation of memory T and B lymphocytes, which can become re-activated upon re-exposure to the antigen.59

It has been demonstrated that transduced cells and professional APCs present immunogenic epitopes derived from the capsid to cytotoxic CD8+ Tcells via MHC class I.40,45,52,60 Cytotoxic Tcells are then responsible for driving clearance of AAV-transduced cells, causing inflammation in the target organ, and decreasing the duration and efficacy of gene transfer.32,41,61,62 In sync with MHC class I presentation, recognition of capsid-derived epitopes bound to MHC class II on the surface of APCs activates CD4+ T helper cells, which facilitate humoral and cell-mediated immune responses.51

Experience from clinical trials suggests that AAV vector immunogenicity is to some extent dose-dependent,39,63 for which low vector doses are more likely to induce a mild inflammation that can be managed and does not result in total loss of transgene expression39 (Figure2). This is consistent with invitro studies showing dose-dependent levels of capsid antigen being presented by transduced cells onto MHC class I.60,64 Additional factors that can drive vector immunogenicity are less known and can include pre-existing tissue inflammation,65 the use of single- or double-stranded vector genomes,54 and the CpG content of the vector genome44,66 (Figure2).

Immune Responses against the Transgene Product

Several factors contribute to shape the immunogenicity of the transgene product in AAV gene transfer. These can be divided into host-specific factors, associated with the underlying disease or the genetic background of the vector recipient, and vector-specific factors, to group factors related to gene transfer (Table 2). Similar to anti-capsid immune responses, presentation of transgene-derived epitopes to CD8+ and CD4+ T lymphocytes can induce cytotoxic and humoral responses that have a negative impact on transgene stability.56,58,67,68 One key factor determining the level of anti-transgene immune responses is the target organ for gene transfer, which is determined by the combination of the AAV capsid, the vector delivery route, and the tissue specificity of the promoter driving gene expression. In particular, systemic and intramuscular vector administration, with either ubiquitous or muscle-specific promoters, have been shown to be more immunogenic than gene transfer to immune privileged organs, as well as systemic administration with liver-specific promoters69,70 (Table 2).

Table 2

Factors Driving Anti-transgene Immunogenicity

Enhanced Transgene ImmunogenicityReduced Transgene ImmunogenicityReferences
Host-Specific Factors
Underlying mutationnull mutations (CRIM-negative)missense mutations (CRIM-positive)71,72
Disease-specific changes in target tissuepresence of inflammation, immunity against self-protein, immune system alterationshealthy73,74,75
Previous exposure to recombinant proteinnaive patients or patients with inhibitors against recombinant proteinselection of patients with no inhibitors against recombinant protein76,77
Vector-Specific Factors
Route of administrationintramuscularsystemic, immunoprovileged organ72,78
Promoterstrong, constitutive, muscle specificliver-specific79,80,81
Vector genomeself-complementary, CpG rich, dsRNAsingle stranded, CpG low, disruption of ITR promoter activity49,54,58,66
Transgeneintracellular, highly glycosylated, largesecretable, native glycosylation pattern, small81,82, 83, 84
Vector doselow hepatocellular expressionhigh hepatocellular expression85

In the context of AAV trials, anti-transgene immune responses have been documented in only a few instances, mostly in the context of intramuscular delivery of AAV vectors. The first evidence of transgene-specific cytolytic Tcell activation came from a phase I/II trial of intramuscular gene transfer of a microdystrophin transgene.73 In this trial, lack of transgene expression was associated with the development of a CD8+ Tcell responses directed against epitopes mapping within the transgene amino acid sequence. Similarly, in a trial for α-1 antitrypsin (AAT) deficiency, following intramuscular delivery of an AAV vector, cytolytic Tcells were detected in association with decreased transgene expression in one of the participants,86 although the trial showed overall long-term expression of the transgene in most participants.4 More recently, in the context of systemic delivery of an AAV vector for the treatment of X-linked myotubular myopathy, anti-transgene antibodies were detected,87 although with no direct correlation with clinical endpoints (vide infra). Finally, transgene-reactive Tcells were also detected in a phase I/II trial of intracranial delivery of an AAV5 vector to the brain of children affected by mucopolysaccharidosis type IIIB.88

The significance of these early findings and their impact on endpointsof safety and efficacy are not entirely clear. What appears to be consistent is that targeting tissues that are not immune privileged, such as the muscle,89 with gene transfer may pose additional challenges related to transgene immunogenicity. To this end, concomitant expression of a transgene in a protolerogenic tissue such as the liver, at the same time as muscle, may help transgene engraftment.69,79

The Tolerogenic Potential of Liver-Directed Gene Transfer

In the context of liver-directed gene transfer, transgene immunogenicity appears to be less of a potential concern compared to other tissues. Starting with the initial observation that mice expressing human factor IX (F.IX) in liver were immunologically tolerant tothe transgene product,80 several studies with AAV vectors in small and large animal models of genetic diseases show that expression ofan antigen in hepatocytes can promote robust antigen-specific immune tolerance (Figure3).80, 90, 91

AAV Vector Immunogenicity in Humans: A Long Journey to Successful Gene Transfer (5)

Liver Gene Transfer Can Drive Transgene Immune Tolerance

Tolerance to a variety of transgenes expressed in the liver is mediated by a variety of mechanisms. Tregs are a common denominator of liver tolerance, as they mediate the suppression of both humoral and Tcell-mediated transgene immune responses. Additional mechanisms include anergy, exhaustion, and deletion of reactive Tcells. Key to tolerance induction appears to be a robust transgene expression in hepatocytes. Adapted from Sherman etal.92

Although several laboratories have investigated the mechanisms driving liver tolerance, results are not fully overlapping and dependon the experimental setting and model antigen used. The tolerogenic effect of liver gene transfer is likely to be mediated by the different cell types that can act as APCs in the liver. Reports indicate that liver-resident macrophages (Kupffer cells) and DCs have a less mature phenotype compared to professional APCs foundin the periphery.93,94 This property seems to make them poor Tcell activators95 as antigens are presented in the absenceof sufficient co-stimulator ligands. Furthermore, it has beenshown that Kupffer cells can secrete IL-10, an anti-inflammatory cytokine, upon TLR stimulation.96, 97, 98 Another particularity is that LSECs areable to act as professional APCs presenting antigens through MHC class II, which seems to play a role in the induction of regulatory Tcells (Tregs).99,100 Several studies have demonstrated the essential role of Tregs in the induction of liver-mediated toleranceto liver-targeted transgenes (Figure3).69,90,91,101 In these studies, disruption of Treg homeostasis around the time of vector administration led to an immune response against the transgene. Conversely, administration of rapamycin, a drug known to favor Treg expansion,102 enhanced efficiency of induction of tolerance in the context of established immunity.103,104

Tregs are not the only player involved in livertolerance. CD4+ Tcell anergy,105 apoptosis of reactive Tcells,80,106 degradation of Tcells in hepatocytes,107 induction of CD8+ Tregs,108 and the acquisition of an exhausted phenotype by cytotoxic CD8+T cells69,85, 109, 110 have also been described in the contextof hepatocellular antigen presentation (Figure3). Inductionoflivertolerance has been shown to be dose-dependent,111 ashigh levels of antigen presentation by hepatocytes through MHCclass I have been associated with a more efficient induction of CD8+ Tcellexhaustion and apoptosis,85,112 although in the contextof AAV gene transfer this has not been fully demonstrated.

While preclinical data on liver tolerance are convincing, with compelling data on immunity eradication in small and large animal modelsof hemophilia76,77,111 and other diseases,79,113 the open question is whether this concept will translate to humans carrying pre-existing immunity against a given therapeutic transgene.

Immune Responses to AAV Vectors in Clinical Trials

Recombinant AAV vectors are relatively simple from an immunogenicity point of view, as they do not encode viral proteins and theycomprise a protein capsid and a DNA genome, which can be single or double stranded. Pre-existing immunity originating fromthe exposure to WT AAV, however, can generate both humoral and cell-mediated immunity to the virus, which can cross-react with AAV vectors. While pre-existing humoral immunity represents one of the most efficient barriers to prevent successful gene transfer through systemic administration of AAV vectors (Figure4),114 theimpact of pre-existing Tcell immunity to AAV is not fully understood.

AAV Vector Immunogenicity in Humans: A Long Journey to Successful Gene Transfer (6)

Humoral Immune Responses to AAV

Pre-existing immunity to AAV can block target tissue transduction when the vector is administered systemically directly into the bloodstream. While eradication of humoral immunity with immunosuppressive drugs can be challenging, as pharmacological targeting of B cells has inherent risks, pre-clinical evaluation of physical removal of antibodies with plasmapheresis has shown promising results. Isolation of target organs at the time of vector administration has also been explored. Recent data linked the acute toxicities observed following systemic administration of AAV vectors with complement activations. These toxicities may be mediated by anti-AAV antibodies and can be modulated by drugs targeting the complement activation pathways. Finally, after gene transfer, antibodies to AAV are induced and persist for the long term. Several potential approaches to vector readministration have been explored preclinically, with variable degrees of success.

Impact and Relevance of Anti-AAV Antibodies

Whereas cytotoxic responses can mostly be controlled by transient immunosuppression,5,63 the elevated NAb titers induced by AAV administration prevent vector readministration, particularly upon systemic delivery (Figure4). This is a potential cause of concern when treating infants,115 since transgene expression following AAVgene transfer is expected to decrease due to tissue growth and dilution of the vector genome.116,117 Furthermore, pre-existing immunity has been shown to prevent cell transduction by AAV vectors,31,32,118 for which the presence of NAbs above a certain threshold is currently an exclusion criterion for the enrollment of patients in AAV-mediated gene therapy clinical studies. Thus,humoral responses against the AAV capsid are still a major hindrance to the clinical application of this invivo approach, although solutions to the problem are being tested in preclinical models.114,119

Ocular Gene Transfer

The eye is a highly compartmentalized organ whose local environment is anatomically isolated from the peripheral immunity by theblood-ocular vasculature, and by the absence of lymphatic vessels.120,121 Therefore, in the context of gene transfer, the eye hasbeen long thought to be at lower risk for activating innate and adaptive immune responses upon vector administration. Moreover, a phenomenon known as anterior chamber-associated immune deviation (ACAID) has been described in the eye upon recognition of immunogenic antigens, which involves the induction of Tregs, anti-inflammatory M2 macrophages, and the generation of an anti-inflammatory cytokine environment that promotes immunological tolerance.122,123 This pro-tolerogenic environment is a protective adaptation aimed at preventing inflammatory responses that could affect vision. An immune deviation phenomenon similar to ACAID has been reported in the context of AAV vector administration to the subretinal space.124

The well-characterized immune privilege, the fact that circulating antibodies directed against the AAV capsid are not normally found in the eye,125 and the relatively low vector doses required to achieve therapeutic efficacy have driven the early successes of ocular gene transfer in the clinic.126,127 Based on these promising results, and banking on more than two decades of research in the field of oculargene transfer, in recent years the field has experienced a dramatic expansion, with several gene transfer trials planned and ongoing, mostly based on the AAV vector platform.128

To date, AAV vector administration has been mainly performed intravitreally or subretinally (Figure5). Based on preclinical animal models and clinical trials, immunogenicity outcomes are greatly influenced by the route of vector administration to the eye, with subretinal vector delivery being less immunogenic than intravitreal administration.129 However, depending on the total vector dose administered, inflammatory responses can be detected regardless of the route of vector administration (vide infra).

AAV Vector Immunogenicity in Humans: A Long Journey to Successful Gene Transfer (7)

AAV Vector Administration to the Eye

Two main routes of vector administration to the eye have been explored. Subretinal administration has been tested in several trials and has a demonstrated safety and efficacy profile in humans. While the approach appears to be feasible, safe, and associated with a low vector immunogenicity profile, it required a surgical procedure that is relatively invasive. Conversely, intravitreal vector administration requires a simple procedure for vector delivery. This route of administration seems to result in higher vector immunogenicity, resulting in inflammation after vector delivery. The safety and efficacy profile of intravitreal delivery of AAV vectors is being evaluated in several trials.

Gene Transfer to the Subretinal Space

Most progress in the subretinal delivery of AAV vectors has been made in the treatment of Leber’s congenital amaurosis type 2 (LCA2) caused by mutations in the retinal pigment epithelium-specific 65-kDa protein (RPE65) gene. So far, all reported studies havebeen based on AAV2 vectors,126,127,130 except for one study in which an AAV4 was administered.131 Herein, we discuss mainly the results from the LCA2 trials, which share similarities in the design of the RPE65 expression cassette and route of vector administration.

In a study sponsored by the University College London (ClinicalTrials.gov: NCT00643747), three young adult participants were injected subretinally with an AAV2 vector carrying the RPE65 transgene under the control of the endogenous RPE65 promoter, at the dose of 1× 1011 vector genomes (vg)126 (Table 3). Patients received tapering immunosuppressive treatment with oral prednisolone for 5weeks, starting 1week prior to vector administration. No capsid- or transgene-specific immune responses were detected.126 After this first report, nine more participants were included in the study and treated at 1× 1011 or 1× 1012 vg, with data of 3 years follow-up published in 2015132 (Table 3). Overall, transient improvements in retinal sensitivity were observed in 6 participants at 6–12months post-injection, but they appeared to decline over time. Immune-related events occurred in five out of eight participants from the high-dose cohort. Two of them developed detectable anti-AAV2 NAbs, and one was also positive for IFNγ responses to the capsid measured by ELISPOT at 4weeks post-injection. In the participant with a positive ELISPOT, inflammation was associated with a worsening in visual function in the injected eye. No immune responses to the transgene were detected.

Table 3

Overview of Clinical Studies Based on Sub-retinal Delivery for the Treatment of LCA2

SponsorPhaseStudy ID: ClinicalTrials.govProductDose (vg)Preventive ISImmune ResponsesReferences
University College LondonI/IINCT00643747AAV2-RPE65-RPE651× 1011 (n= 4)yesNAbs in one participant; inflammatory responses in five participants126,132
1× 1012 (n= 8)
University of PennsylvaniaINCT00481546AAV2-CBSB-RPE655.96× 1010 (n= 3)noanti-capsid pre-existing humoral and cellular immunity; humoral response in four participants; cellular response in one participant130,133
8.94× 1010 (n= 3)
11.92× 1010 (n= 3)
17.88× 1010 (n= 5)
7.95x1010 (n= 1)
Spark TherapeuticsINCT00516477AAV2-CB-hRPE65v2 (voretigene neparvovec)1.5× 1010 (n= 3)yesinduction of anti-AAV2 NAbs in six participants; cellular responses in two participants127,134,135
4.8× 1010 (n= 6)
1.5× 1011 (n= 3)
Spark TherapeuticsINCT01208389AAV2-CB-hRPE65v2 (voretigene neparvovec)1.5× 1011 in the contralateral eye (n= 11)yesno significant immune responses2,136
Spark TherapeuticsIIINCT00999609AAV2-CB-hRPE65v2 (voretigene neparvovec)1.5× 1011 in both eyes (n= 29)yesmild ocular inflammation in two participants137,138
Applied Genetic TechnologiesI/IINCT00749957AAV2-CBSB-RPE651.8× 1011 (n= 6)noocular inflammation in two participants139,140
6× 1011 (n= 6)
Nantes University HospitalI/IINCT01496040AAV4-RPE65-RPE651.22× 1010 (n= 1)yesanti-AAV4 IgG in three patients; ocular inflammation in three patients; cellular immune responses against capsid and transgene in one participant131
1.82× 1010 (n= 2)
3.23× 1010 (n= 1)
4.27× 1010 (n= 2)
4.7–4.8× 1010 (n= 3)

IS, immunosuppression.

In a study conducted by the University of Pennsylvania (ClinicalTrials.gov: NCT00481546), three participants initially received asingle subretinal injection of an AAV2 at 5.96× 1010 vg, in with the expression of the RPE65 transgene was regulated bytheconstitutive chicken beta actin (CB) promoter containing an optimized cytomegalovirus (CMV) enhancer (AAV2-CBSB-hRPE65) (Table 3).130,141 In this study no systemic immunosuppression was administered, and a mild increase in humoral and cellular responses to the AAV2 capsid was detected in one participant. Twelve more subjects were treated with this vector in a dose-escalation study(Table 3).133 Most participants presented anti-AAV2 antibodies prior to vector administration, and seven patients hadpre-existing AAV2-specific Tcells at baseline. Nevertheless, only 4 out of 15 subjects experienced a significant increase ofcirculating anti-AAV2 antibody titers, and one had a positive IFNγ ELISPOT on day 90, possibly highlighting the limited potential of antigens delivered to the eye to stimulate peripheral memory responses. Administration of two vector boluses in thesame eye did not appear to result in enhanced immunogenicity. Despite the promising results in visual sensitivity, long-term evaluation performed after 5–6 years of treatment showed that visualfunction tended to peak between 1 and 3 years, and declined afterward, in some subjects returning to baseline.142

In the work initially sponsored by the Children’s Hospital of Philadelphia and later by Spark Therapeutics (ClinicalTrials.gov: NCT00516477), the first report from 2008 involved three young adultLCA2 patients treated by subretinal injection with an AAV2-CB-hRPE65v2 vector containing an optimized Kozac sequence, at 1.5× 1010 vg in one eye, with concomitant local and systemic corticosteroid immunosuppressive treatment (Table 3).127 No significant immune responses were detected against capsid or transgene, except for a transient increase in anti-AAV2 antibodies in one participant. Twelve additional participants were included in higher dose cohorts of 4.8× 1010 and 1.5× 1011 vg.134, 135, 143 Two out of 3 participants from the high-dose cohort had mild and transient activation of capsid-specific Tcell responses detectable in peripheral blood, and 6 out of 12 had a transient increase in anti-AAV2 NAbs.Ina follow-up study, 11 previously treated patients receiveda second vector injection in the contralateral eye (ClinicalTrials.gov: NCT01208389). All subjects were injected with the highestvector dose in the second eye (1.5× 1011 vg), between 1.7 and 4.58 years after the first injection (Table 3). Results showed no significant induction of cellular or humoral memory responses against the vector or transgene induced after vector readministration, further confirming the low immunogenicity profile of subretinal delivery of AAV vectors.2,136

A phase III study was later initiated (ClinicalTrials.gov: NCT00999609), and 29 subjects were treated subretinally in both eyes at 1.5× 1011 vg (Table 3).137,138 All participants received oral treatment with prednisolone starting prior to the first injection andcontinuing until the time of the second injection, which took place 6–18days after the first procedure. No signs of significant immune responses were documented, except for a mild eye inflammation observed in two participants. The treatment was recently approved by the US Food and Drug Administration (FDA) and European Medicines Agency (EMA). The drug product is marketed as Luxturna and manufactured by Spark Therapeutics.144 A long term-follow up study is currently open (ClinicalTrials.gov: NCT03597399).

Additional trials have been conducted for the treatment of LCA2based on AAV2 (ClinicalTrials.gov: NCT00749957 and NCT00821340),139,140 AAV4 (ClinicalTrials.gov: NCT01496040),131 or AAV5 (ClinicalTrials.gov: NCT02946879 and NCT02781480) vectors encoding RPE65.128,145 Among the studies with recently published results, one study sponsored by the Nantes University Hospital was based on AAV4,131 which presents specific tropism for RPE cells (Table 3). The transgene was expressed under the endogenous RPE65 promoter, and the vector was administered in multiple injections in one eye in the absence of immunosuppression. Anti-AAV4 antibodies were observed in three out of nine patients, three of them also experienced transient inflammation, and one patient showed detectable IFNy responses against both capsid and transgene.131

Several additional clinical studies of subretinal delivery of AAV vectors are ongoing for indications that include choroideremia,146,147 X-linked retinitis pigmentosa,148 neovascular (“wet”) age-related macular degeneration (wAMD),149, 150, 151 and others.128 Overall, these studies evidence that immune responses against AAV occur in a fraction of patients despite the immune privilege of the eye and the use of immunosuppression. Anti-transgene immune responses are rare, but they were also reported in one study.131

Currently, administration of corticosteroids is broadly used as a measure to limit both the inflammation derived from the administration procedure and to modulate vector immunogenicity. Despite this, instances of inflammation have been observed across trials, and particularly as higher doses of vector were administered. Potential solutions have been proposed to decrease vector immunogenicity,152 and it is likely that more effective solutions than corticosteroid administration will be needed to tackle gene therapy strategies requiring high vector doses, such as those relying on dual AAV vectors.153 Similar considerations are also likely to apply to the fieldof gene editing, as they often rely on AAV vectors as vehicles for donor templates154, 155, 156, 157 and because of the potential issues related to the immunogenicity of Cas9.158,159

Intravitreal AAV administration

Intravitreal vector administration has been preclinically explored fora variety of indications, based on the ease of vector administration through this route (Figure5) and on the need, for some indications, totarget specific cell types.160,161 The potential for some AAV vectorserotypes to reach the outer layers of the retina162 has further heightened the interest on this route of vector administration.

Intravitreal vector administration has been explored in in the context of few clinical trials. Wan etal.161 evaluated the safety and efficacy of an AAV2 vector encoding for mitochondrial reduced nicotinamide adenine dinucleotide (NADH) dehydrogenase 4 gene (ND4) administered to the vitreous in nine Leber’s hereditary optic neuropathy (LHON) patients (ClinicalTrials.gov: NCT01267422). A 9-week course of prednisolone treatment was given around vector administration and subjects with high-titer, pre-existing NAbs were excluded. No significant humoral or cellular immune responses against the vector were documented in this study, which also showed no evidence of therapeutic efficacy. In another study, unilateral intravitreal injection of aself-complementary AAV2-ND4 vector was performed in 14LHON patients, in the absence of immunosuppression (ClinicalTrials.gov: NCT02161380).160,163 Neutralizing antibodies to AAV2 were detected in all participants prior to injection, and increased afterward at variable titers, and in two subjects development of high antibody titers were associated with anterior uveitis. Finally, in a phase I/II study sponsored by GenSight Biologics (ClinicalTrials.gov: NCT02064569), an AAV2-ND4 vector was administered by a single intravitreal injection at increasing doses in the absence of corticosteroids.164,165 Most participants experienced inflammation in the anterior chamber and vitreous, which has not been directly associated with the vector dose, and two participants were given oral corticosteroids to modulate the response. Levels ofanti-capsid antibodies increased after injection in most patients enrolled in the trial.164,165 Follow-up of LHON patients enrolled ina phase III trial is currently ongoing. In this study a bilateral injection of the vector at 9× 1010 vg/eye was performed (ClinicalTrials.gov: NCT03293524).

Additional studies of intravitreal gene transfer of AAV vectors wereconducted in the context of wAMD. Heier etal.166 used an AAV2 to deliver the soluble vascular endothelial growth factor (VEGF) receptor sFlt-1 (AAV2-sFLT01) at doses between 2× 108 and 2× 1010 vg in 19 subjects (ClinicalTrials.gov: NCT01024998). Two patients in the high-dose cohort 4 experienced pyrexia and intraocular inflammation upon vector administration, which was treated with topical steroids. Transgene expression was detected in 5 out of 10 subjects who had low or no pre-existing NAbs, which peaked on week 26 and declined during the 52-week follow-up period. The lack of transgene expression in the remaining patients correlated with the presence of high antibody titers against AAV2 prior to vector administration.166 Similarly, in an ongoing trial forwAMD sponsored by Adverum Biotechnologies, vector doses up to 6× 1011 vg were tested (ClinicalTrials.gov: NCT03748784) along with a tapering course of corticosteroids. Instances of ocular inflammation were observed at the highest vector dose tested, prompting the extension of the cohort receiving a vector dose of 2× 1011 vg.

Similar to the subretinal delivery of AAV vectors, inflammation caused by intravitreal delivery of AAV vectors appears to be dose-dependent. Thus, similar considerations apply to the need to better understand the determinants of vector immunogenicity in this context, to allow the development of strategies to overcome inflammation. At a minimum, manufacturing technologies that allow the elimination of or greatly reduce the content of non-infectious vectors particles (i.e., empty capsids) would likely be beneficial.

Liver Gene Transfer

The liver has been a long-suited target for gene transfer, as research efforts have been directed to targeting the liver for the correction of a variety of genetic and metabolic diseases167, 168, 169, 170, 171 and for other applications in which, for example, hepatocytes are turned into biofactories to supply protein therapeutics directly into the bloodstream.5,81, 172, 173, 174 Several important features make the liver an ideal organ for gene therapy, including (1) the fact that hepatocytes arecentral to several metabolic functions and secrete a variety of proteins into the circulation; (2) its high degree of vascularization, which allows for easy transduction with any gene therapy vector delivered thorough the bloodstream; and (3) its unique immune protolerogenic environment (vide supra).

The first evidence of immune responses in the context of liver genetransfer was reported in a trial conducted by the Children’s Hospital of Philadelphia and Avigen, in which seven subjects with severe hemophilia B received a single-stranded AAV2 vector carrying the F.IX transgene under the control of a liver-specific promoter, through the hepatic artery (Table 4).32 Patients were divided intothree dose cohorts of 8× 1010, 4× 1011, and 2× 1012 vg/kg, respectively. Transient elevation of F.IX activity at levels around 10% of normal was detected only in the first patient treated at the highest dose. Differently from animals models, in which long-term expression is observed after liver gene transfer with AAV vectors,41,175 4weeks after gene transfer, F.IX expression started to decline and eventually returned to pre-treatment level. A self-limited transient and asymptomatic rise in liver transaminases was also detected in the same time frame, suggesting the induction of cytotoxic immunity against the AAV2 capsid.32 Similar series of events were observed in aparticipant enrolled in the mid-dose cohort (subject G). An IFNγ ELISPOT assay and flow cytometry-based detection of AAV-specific CD8+ Tcells confirmed a Tcell response to the AAV2 capsid, which had a kinetics of appearance that closely mirrored the rise in serumtransaminases and decrease in transgene expression.32,40 Anadditional important finding from this trial was that a second subject dosed with 2× 1012 vg/kg, a dose expected to result in detectable transgene expression, but carrying pre-existing anti-AAVNAbs at a titer of just 1:17, had no evidence of liver expression.32

Table 4

Overview of Hemophilia B Clinical Studies Based on Liver-Directed Gene Transfer

SponsorPhaseStudy ID: ClinicalTrials.govProductDose (vg/kg)Immune ResponsesReferences
AvigenI/IINCT00076557ssAAV2-LP1-FIX8× 1010 (n= 2)pre-existing NAbs; dose-dependent cellular responses to the capsid and elevation of liver transaminases32
4× 1011 (n= 3)
2× 1012 (n= 2)
St. Jude Children’s Research HospitalINCT00979238scAAV8-LP1-FIXco2× 1011 (n= 2)anti-capsid cellular responses; elevation of liver transaminases in four patients resolved with prednisolone5,63
6× 1011 (n= 2)
2× 1012 (n= 6)
Baxalta/ ShireI/IINCT01687608scAAV8-TTR-FIXR338Lopt (AskBio009, BAX 335)2× 1011 (n= 2)dose-dependent Tcell responses and elevation of liver transaminases in two patients resolved with prednisolone176,177
1× 1012 (n= 4)
3× 1012 (n= 2)
UniQureI/IINCT02396342scAAV5-hAAT-FIX (AMT-060)5× 1012 (n= 5)elevation of liver transaminases in three patients resolved with prednisolone178
2× 1013 (n= 5)
UniQureIIbNCT03489291scAAV5-hAAT-FIXPadua (AMT-061, etranacogene dezaparvovec)2× 1013 (n= 3)pre-existing NAbs; no significant inflammatory responses179
PfizerI/IINCT02484092SPK-90015× 1011 (n= 15)elevation of liver transaminases in two patients resolved with prednisolone62,180
Dimension Therapeu-ticsI/IINCT02618915ssAAVrh10-FIX (DTX101)1.6× 1012 (n= 3)elevation of liver transaminases treated with prednisolone181
5× 1012 (n= 3)

Based on the learning from this first trial, and on preclinical datashowing the potential superiority of the AAV8 serotype in transducing the hepatocytes,182 a second clinical study was initiated a few years later (ClinicalTrials.gov: NCT00979238). The vector genome carried a self-complementary genome, a strategy to enhance hepatocyte transduction.183 In addition, the F.IX sequence was codon optimized to further improve transgene expression. Six seronegative participants with severe hemophilia B were infused an AAV8 vector encoding the self-complementary, codon-optimized, liver-targeted F.IX transgene in a peripheral vein.63 In the low- and intermediate-dose cohorts, F.IX activity remained between 1% and 4% of normal, respectively, in the absence of immune-associated events. In the highest dose cohort (2× 1012 vg/kg), however, F.IX activity reached maximum levelsof 8%–10% of normal, and patients again experienced an elevation of liver enzymes at 8weeks, which coincided with partialdecline in transgene expression and a rise in circulating capsid-specific Tcells detected by an IFNγ ELISPOT assay. Theelevation of liver enzymes was resolved with tapered treatmentwith high-dose steroids, after which transgene expression remained stable.63 Four more patients were enrolled in the high-dose cohort, two of which were transiently treated with prednisolone after experiencing a decline in F.IX expression,5 afterwhich transgene expression remained stable with a documented duration of up to 10 years (U.M. Reiss, 2019, National Hemophilia Foundation, conference).

Although effective at the currently used vector doses, ongoing studies will address whether this corticosteroid regimen will be effective at higher vector doses and with different AAV serotypes. Of note, evidence of Tcell activation in response to AAV capsid epitopes was documented in PBMCs isolated from participants from all dose cohorts, although liver toxicity was documented only in the high-dose cohort. High levels of anti-AAV8 antibodies could be detected in all participants following vector administration.

A similar set of clinical results was also recently published in thecontext of another hemophilia B clinical trial in which a bioengineered AAV capsid was administered to seronegative hemophilia B subjects at a dose of 5× 1011 vg/kg. At this dose, 2 out of 10 participants required the administration of corticosteroids to managevector immunogenicity, resulting in long-term transgene expression in all subjects enrolled.62

Corticosteroids given reactively have been broadly used across trialsto modulate capsid reactivity. However, this measure did not succeed in all studies in rescuing transgene expression. In a phase I/II study initially sponsored by Baxalta/Shire (ClinicalTrials.gov: NCT01687608), a self-complementary AAV8 vector176 was given to eight hemophilia B subjects at doses ranging from 2× 1011 to 3× 1012 vg/kg. Transgene expression was transient in most subjects in this study, with the exception of one subject who did not experience an instance of liver enzyme elevation and expressed the F.IX transgene over the long term. Exome sequencing analysis in this subject revealed the presence of a polymorphism potentially affecting the immune system, which could explain the lack of immune responses detected after gene transfer.184 Similarly, a phase I/II study sponsored by Dimension Therapeutics was halted due to transaminase elevation and loss oftransgene expression not controlled by the administration of high-dose corticosteroids.181 Additional emerging data from other liver trials (ClinicalTrials.gov: NCT02991144, NCT03636438, NCT03517085, NCT03970278, and NCT03223194), and also trials in which high doses of AAV vectors are given systemically for the treatment of diseases not affecting the liver (ClinicalTrials.gov: NCT03368742 and NCT03362502),185,186 show instances of liver toxicities that are likely to be related to vector immunogenicity.

A somewhat different outcome was documented in the context of trials in which AAV5 vectors were used to target the liver. This capsid serotype was first used in subjects affected by acute intermittent porphyria.171 In this study, vector doses up to 1.8× 1013 vg/kg were administered, with no evidence of transaminase elevation and no detection of capsid-specific Tcells, although also no clear evidence of therapeutic efficacy. Two additional studies were published, in which AAV5 vectors produced in a baculovirus system were infused at doses up to 5× 1012 vg/kg178 and 6× 1013 vg/kg187 for the treatment of hemophilia B and A, respectively. In these two studies, elevated transaminase levels were detected and treated with corticosteroids, although their direct impact on transgene expression was not clearly established in the time frame of the initial observation period. Longer follow-up of participants in the hemophilia A trial showed a slow decline of expression during a few years after gene transfer, for which the exact etiology remains unknown.

AAV Vectors and Liver Toxicities: Lessons Learned and Open Questions

Enormous progress has been made from the first liver gene transfer trial for hemophilia B. A deeper understanding of some of the factors driving immune responses has provided tools to modulate AAV vector immunogenicity. Many lessons learned, with some important open questions, include the following:

  • It is now well accepted that AAV capsid can trigger dose-dependent immune toxicities that can limit the duration of transgene expression in hepatocytes. Some exceptions remain, such as that of AAV5, although the fact that this serotype is known to be less efficient than others in transducing hepatocytes may explain the higher threshold for Tcell activation.

  • Vector design has an equally important role as the vector dose as a determinant of immunogenicity. Emerging data from clinical trials and preclinical models44,58 indicate, for example, that the content of CpG in the vector genome greatly increases vector immunogenicity.

  • The liver is the first-line target organ when AAVs are administered systemically. Thus, immune-mediated liver toxicities can be encountered any time an AAV vector is given in a high dose and systemically.186

  • Long-term duration of liver-mediated expression for up to 10 years has been demonstrated in the context of hemophilia B trials. For other liver gene transfer applications, such as hemophilia A, data are still being gathered,188 and the factors driving vector persistence are not entirely understood.

  • Different manufacturing methods18 result in vector preparations with impurities at levels that can vary in nature and quantity. Some of these impurities can play a role as immunogenicity determinants, including excess of non-infectious capsids, DNA contaminants,189 or potential protein carryover.190

  • Finally, the role of pre-existing immunity to AAV needs to be better understood in its potential role as a determinant of vector immunogenicity.

Muscle Gene Transfer

Muscle is an important target tissue for gene transfer to treat a variety of neuromuscular,191 cardiac,192 and metabolic diseases.193 Moreover, even for diseases not affecting the muscle, targeting the muscle withAAV vectors has been shown to yield promising results.194, 195, 196 In this section we focus on muscle gene transfer. Diseases mostly affecting motor neurons are discussed separately.

From an immunological perspective, the muscle has unique challenges (reviewed in detail by Boisgerault and Mingozzi89). For example, the inflammation associated with dystrophies can result in enhanced transgene immunogenicity.73 Conversely, intramuscular gene transfer has been shown to induce the recruitment of FoxP3+ Tregs, which may play a role in contributing to the stability of transgene expression.56,197 Additional challenges related to targeting the muscle include the limited distribution of AAV vectors delivered intramuscularly198,199 and the need for high vector doses to achieve widespread transduction of skeletal muscle across the entire body.

Intramuscular Delivery of AAV Vectors

The first trial in which an AAV vector was injected intramuscularly inhumans was a study in hemophilia B (Table 5)118,200 in which eightadults received an AAV2 vector encoding for F.IX intramuscularly at doses up to 1.8× 1012 vg/kg. Administration of the vector waswell tolerated with no documented immune-related toxicities (although no specific assays were used to measure Tcell responses to the AAV capsid) and no evidence of vector neutralization in participants with pre-existing NAbs to AAV2. Transgene expression was persistent for the long term,3 although it failed to reach therapeutic levels,118,200 highlighting the challenge of administering highandtherapeutically relevant doses of AAV vector intramuscularly while limiting the risk of induction of local immune responses.78,118,200

Table 5

Overview of Clinical Studies Based on Intramuscular AAV Injection

DiseaseSponsorPhaseStudy ID: ClinicalTrials.govProductDoseImmune ResponsesReferences
Hemophilia BAvigenI/IINCT00076557ssAAV2-CMV-FIX2× 1011 vg/kg (n= 3)pre-existing NAbs; dose-dependent humoral responses to AAV118
6× 1011 vg/kg (n= 3)
1.8× 1012 vg/kg (n= 2)
ATT deficiencyUniversity of MassachusettsINCT00377416AAV2-CB-AAT2.1× 1012 vg (n= 3)pre-existing Nabs201
6.9× 1012 vg (n= 3)
2.1× 1013 vg (n= 3)
6.9× 1013 vg (n= 3)
University of MassachusettsINCT00430768AAV1-CB-AAT6.9× 1012 vg (n= 3)pre-existing NAbs, increase after injection; anti-capsid cellular responses in all evaluated participants202
2.2× 1013 vg (n= 3)
6.0× 1013 vg (n= 3)
Applied Genetic TechnologiesIINCT01054339AAV1-CB-AAT6.0× 1011 vg/kg (n= 3)pre-existing NAbs; cellular responses to the capsid in all participants; cellular responses to the transgene in two participants; detection of Tregs and exhausted Tcells4,86,203
1.9× 1012 vg/kg (n= 3)
6.0× 1012 vg/kg (n= 3)
LPL deficiencyAmsterdam Molecular TherapeuticsIICT-AMT-010-01AAV1-LPLS447X1× 1011 vg/kg (n= 4)pre-existing NAbs; cellular responses to the capsid in four participants39,204
3× 1011 vg/kg (n= 4)
Amsterdam Molecular TherapeuticsIINCT01109498 (CT-AMT-011-01)AAV1-LPLS447X (alipogene tiparvovec)3× 1011 vg/kg (n= 2)pre-existing NAbs; anti-capsid cellular responses in nine participants205
3× 1011 vg/kg+ IS (n= 4)
1× 1012 vg/kg+ IS (n= 8)
Amsterdam Molecular TherapeuticsII/IIINCT00891306 (CT-AMT-011-02)AAV1-LPLS447X1× 1012 vg/kg+ IS (n= 5)pre-existing NAbs; anti-capsid cellular response in all participants; detection of Tregs and exhausted Tcells206,207,208
DMDNationwide Children’s HospitalINCT00428935AAV2.5-CMV-minidystrophin2.0× 1010 vg/kg (n= 3)cellular responses to the transgene in four participants73
1.0× 1011 vg/kg (n= 3)

IS, immunosuppression.

Similar results were obtained in clinical trials for AAT deficiency andlipoprotein lipase (LPL) deficiency, in which vectors encoding the deficient enzyme where administered intramuscularly (Table 5). In the gene therapy trials for AAT deficiency, a mild increase in transgene expression levels was documented and persisted for the long term,201,202 at least in subjects enrolled in the highest vector dose cohort (6× 1013 vg).202 AAV1 vector administration was associated with the development of anti-capsid Tcell responses in all evaluated subjects.202 Flotte etal.203 reported a dose-dependent increase in AAT transgene levels in the circulation and a transient increase in creatinine kinase (CK) levels in most patients from the higher dose cohorts, and all patients developed humoral and Tcell responses against the capsid. Although no anti-AAT antibodies were detected, cellular responses against the transgene product weredocumented, although the significance of the findings is unclear.86,203 Despite the presence of cellular responses, transgene expression persisted at low levels up to 5 years in one patient from the highest dose cohort, leading to partial disease correction in that patient.4 Persistent transgene expression seemed to be associated with the recruitment of Tregs into the muscle, and to the exhaustion of effector Tcells that expressed the inhibitory molecules PD-1 and PD-L1.4,195,203

In the context of LPL deficiency trials, a first study reported the intramuscular administration of an AAV1 vector produced in HEK293 cells,18 carrying the sequence of a mutant version of the LPL enzyme (LPLS447X). The vector was injected into eight patients divided in two dose cohorts of 1× 1011 and 3× 1011 vg/kg.204 Despite the initial dose-dependent increase in LPL expression, the efficacy of gene transfer decreased in the long term with return to baseline levels after18–31months. The authors associated this effect with the presence of dose-dependent Tcell responses against the capsid, which correlated with higher levels of anti-AAV1 IgG3 subclass antibodies, andwith a transient increase in CK levels in plasma.39,204 No antibody responses against the transgene product were detected. In view of these results, two new studies tested the efficacy and anti-capsid immune responses when administering the same vector, this time produced in a baculovirus system,18 together with immunosuppression (Table 5).205, 206, 207 All patients developed detectable cellular and humoral responses against the AAV capsid. Nevertheless, in the LPL trials206,207,208 sustained transgene expression during the follow-up period was reported, together with the detection of Tregs and Tcells with an exhausted phenotype, as reported in the AAT deficiency clinical trials4,195 and in preclinical studies.209

These studies provide evidence that direct intramuscular delivery can elicit both inflammatory and tolerogenic responses in the local environment where the vector is delivered. This has also been elegantly modeled in mouse studies in which the role of Tregs in the regeneration of injured muscle was demonstrated,210 possibly highlighting unique features of skeletal muscle when it comes to inflammation.

A different outcome was documented in a gene transfer trial for duch*enne muscular dystrophy (DMD).73 In this study, lack of transgene expression in muscle was attributed to the expansion of cytotoxic Tcells specific against epitopes found within the transgene product. Further studies showed the presence of preexisting CD4+ and CD8+ Tcell-mediated immunity against the transgene product, which was detectable in 52.9% of DMD patients naive to gene transfer and was likely due to spontaneous dystrophin expression in revertant fibers (Table 5).73,74 The prevalence of anti-dystrophin immunity was decreased to 20.8% in patients receiving corticosteroid treatment.74

Based on these studies, it can be extrapolated that several factors can shape the immune responses in intramuscular gene transfer,89 particularly when it comes to transgene-driven immunity. Furthermore, targeting all muscle groups in the body at therapeutically relevant doses is limited by this delivery route.

Intravascular and Systemic Delivery of AAV Vectors

Based on the early experience with intramuscular gene transfer, and on the need to target virtually the entire body to achieve therapeutic efficacy in many systemic neuromuscular disorders, research efforts in relatively recent years have focused on the systemic delivery of AAV vectors. Early studies in large animal models of hemophilia B and DMD showed that the intravascular delivery of AAV vectors was less immunogenic compared to intramuscular delivery, even in the absence of immunosuppression,211 owing to the more evenly distributed transduction of muscle fibers, and possibly due to some levels of leaky tolerogenic expression of the transgene in the liver.69,109

Progress in AAV vector large-scale manufacturing helped to spearhead a number of gene therapy trials in which large vector doses were administered systemically to transduce the entire body. Among these, Audentes Therapeutics initiated a phase I/II study in infantile X-linked myotubular myopathy patients, consisting of systemic administration of an AAV8 vector coding for the MTM1 gene under the control of a muscle-specific promoter given at doses up to 3× 1014 vg/kg (ClinicalTrials.gov: NCT03199469).185 Vector administration was well tolerated, and evidence of therapeutic benefit was documented. From an immunogenicity perspective, both transgene and capsid immune responses were detected, along with an increase in liver enzymes and in some cases increased creatinine kinase andtroponin, possibly linked to some degree of systemic immune-mediated toxicity. All subjects enrolled in this trial received prophylactic corticosteroids.

Three additional studies of systemic AAV vector delivery for DMD currently ongoing are also worth mentioning because of specific challenges encountered. In at least two of the three trials, one sponsored by Solid Biosciences (ClinicalTrials.gov: NCT03368742) and another by Pfizer (ClinicalTrials.gov: NCT03362502), acute toxicities were experienced, possibly related to complement activation. There were no reports of such toxicities in a similar trial sponsored by Sarepta Therapeutics (ClinicalTrials.gov: NCT03375164). Also in these trials, large doses of AAV vectors were administered (in excess of 1014 vg/kg), and prophylactic corticosteroids were administered. Inan attempt to modulate acute toxicities, the monoclonal antibody eculizumab was administered in some of these trials.

Overall, emerging data from systemic AAV vector delivery at high doses highlight both the remarkable therapeutic potential of the approach and unexpected, possibly immune-related toxicities. As publicly available data are somewhat limited at the time that this article was written, the etiology of the findings is not entirely clear, although they seem to concern multiple AAV serotypes across trialsfor various indications and tissue targets. Future studies and follow-up of subjects treated with high-dose AAV gene transfer will be useful to better understand the potential effect of the toxicities encountered on the safety and long-term efficacy of gene transfer.

Gene Transfer to Motor Neurons and the Central Nervous System

Systemic gene transfer to target motor neurons has been validated inthe context of the clinical development of a gene therapy treatment for spinal muscular atrophy (SMA) that is marketed by AveXis-Novartis as Zolgensma. Efficacy of this approach was been initially shown in 15 infantile patients with mutations on both alleles of the SMN1 gene and two copies of the SMN2 gene. Patients were injected systemically with an AAV9 vector carrying the SMN gene under the control of a ubiquitous promoter.186 Young pediatric patients were divided in two cohorts of 6.7× 1013 and 2.0× 1014 vg/kg. As in other systemic delivery studies, because most of the vector delivered systemically ends up in the liver, elevated liver transaminases were observed. Concomitantly, capsid-specific Tcells were detected in peripheral blood and were attenuated by a course of prednisolone. Overall, four participants enrolled in this study showed transient elevation in transaminases. Based on the groundbreaking clinical results,186,212,213 confirmed in a phase 3 trial (ClinicalTrials.gov: NCT03306277), the investigational gene therapy drug was recently approved.214

While there is little doubt about the potential therapeutic benefit of AAV-mediated gene transfer in the context of systemic neuromuscular diseases and in particular diseases affecting motor neurons, also in this context recent preclinical findings possibly related to vector immunogenicity are worth mentioning as potentially relevant. In a recent report,215 toxicities associated with systemic delivery of high doses of AAV vectors encoding for SMN were observed in non-human primates (NHPs) and piglets. Liver enzyme elevation, including acute liver failure in one animal, was observed in NHPs. Dorsal root ganglia (DRG) sensory neuron lesions were also detected in both species, with a higher degree of severity in piglets. Toxicities described in this study appeared to be independent of capsid cellular immune responses. Similar symptomatic DRG degeneration was also described in a preclinical study conducted by AveXis-Novartis in NHPs following the intrathecal delivery of an AAV9-SMN vector,216 prompting a hold on an ongoing trial of intrathecal vector delivery in SMA type 2 patients (ClinicalTrials.gov: NCT03381729). The nature of these preclinical findings is yet to be fully elucidated. To date, no similar toxicities were encountered in humans receiving the same AAV9-SMN vector systemically.186 Several questions about these findings are still unanswered, including whether the observed toxicities are driven by the vector capsid, the transgene product, or both and whether they are in some way immune mediated. Of note, no symptomatic toxicities have been reported in an ongoing gene transfer trial for giant axonal neuropathy in which high doses of an AAV9 vector are delivered intrathecally together with a course of immunomodulatory drugs (ClinicalTrials.gov: NCT02362438).

In contrast to systemic vector delivery, because of the physical separation to the systemic circulation mediated by the blood-brain barrier (BBB), brain delivery of AAV vectors has several peculiarities, including (1) that the cerebrospinal fluid (CSF) generally contains lower levels of antibodies able to neutralize AAV vectors compared to what is found in the general circulation; and (2) that the brain has a distinct subset of resident immune cells composed mainly of specialized macrophages called microglia, which, because of the lack of access to brain samples, are difficult to study. Immune responses in the human central nervous system (CNS) following gene transfer seem to be less frequent than in other body sites.217 However, the tools we are using to track these responses are likely to be not sensitive enough to detect, for example, local asymptomatic, discrete responses that could lead to gradual clearance of transduced cells.

Pre-existing NAbs to AAV do not appear to be an issue in the context of direct AAV vectors into the CNS. For instance, it has been shown that the presence of circulating anti-AAV-NAbs up to a 1:128 titer had no inhibitory effect on efficacy of the AAV9-GFP gene transfer when delivered to the CSF in NHPs.218 Conversely, the concentration of AAV9-NAbs in circulation increases post-vector infusion, both systemically and in the CSF.219

As mentioned, the brain possesses specialized resident immune cells that are yet to be studied in terms of their response to the AAV vectors. Microglia are initial responders to pathogens or tissue damage and are responsible for initiating an inflammatory response in brain. Alternatively, while phagocytosing dead or dying cells, microglia actually prevent the release of pro-inflammatory signals from necrotic tissue and thus limit further brain damage.220 Additionally, microglia can also phagocytose viable cells that were infected by a virus, as a consequence, for example, of the intracellular calcium dysregulation and phosphatidylserine externalization observed in the context of adenovirus infection.221 However, it is not certain whether transduction of neurons by AAV vector could elicit levels of cellular stress that can lead to clearance of transduced cells by microglia.

The infiltration of microglia to inflammation sites can be followed in tissue sections by fluorescent microscopy detecting ionized calcium binding adaptor molecule 1 (Iba1)-positive cells. The activation statecan be inferred from the cellular morphology, where a resting microglial cell has elongated ramified processes that rapidly retract upon recognition of a pathogen or other inflammatory stimuli. Microglial cells will, in these circ*mstances, become a mobile effector cell.222 Obviously, this type of analysis is not feasible for the purpose of tracking immune responses during clinical trials, which are commonly limited to testing anti-capsid and anti-transgene antibodies and Tcells in peripheral blood or CSF.

Direct delivery of AAV vectors into the brain causes little or no response as measured by IFNγ ELISPOT in PBMCs from infused subjects (reviewed elsewhere223). One important feature to consider in this particular clinical setting is that the doses of AAV vectors administered thus far were relatively small. Nevertheless, a direct injection in brain can also be associated with a local inflammatory response. Similar to muscle, systemic infusion of AAV vectors into the CSF may decrease vector immunogenicity, although, because it requires higher vector doses, it does increase systemic vector exposure. For example, in a mouse model of Niemann-Pick disease type A (NPD-A), it was shown that the delivery of AAV9-ASM to the CSF through the cerebellomedullary (CM) cistern led to transgene expression within the central and peripheral nervous systems. When compared to the direct intracerebellar injection, this route of delivery did not trigger inflammation.224 Similarly, administration of AAV9 vectors into the CSF of dogs resulted in widespread transduction of the brain,219 with also leakage of the vector into the systemic circulation and transduction of the liver.

Across the ongoing trials, the approaches to modulation of potential immunogenicity associated with gene transfer in the CNS vary significantly, ranging from concomitant use of combination of immunosuppressive drugs (ClinicalTrials.gov: NCT02362438, NCT03612869) for an extended period to no immunosuppression (EudraCT: 2015-000359-26). A published example comes from a clinical trial in subjects with Canavan disease with mutations on the gene encoding for aspartoacylase (ASPA) who received an AAV2-ASPA vector infused intracranially.225 No immunosuppression was used during this trial, and only 3 out of 10 subjects developed low to moderately high levels of AAV2 NAbs in blood compared to baseline. In another trial, patients with metachromatic leukodystrophy (MLD) were administered AAVrh.10-arylsulfatase A (ARSA) by intracerebral injection (ClinicalTrials.gov: NCT01801709). A short course of steroids was used to avoid potential immune reactions starting 1day prior to the vector infusion and given for 10days. Anti-capsid antibodies were detected in blood and CSF of all patients. Only one out of four participants developed anti-ARSA antibodies detectable in blood but not in CSF (C.G. Bonnemann, 2019, National Center for Advancing Translational Sciences/NIH, conference). Another example of immune suppression combined with the CNS gene transfer may be illustrated by the AAV9-mediated delivery of gigaxonin for cross-reactive immunologic material (CRIM)-negative giant axonal neuropathy patients (ClinicalTrials.gov: NCT02362438). To mitigate the risk for CRIM-negative participants to mount an immune response against the transgene, an immunosuppressive regimen including overlapping dosing of methylprednisolone, prednisone, tacrolimus, and rapamycin was used. The approach was successful in preventing inflammation, as in immunosuppressed participants no pleocytosis was detected. After vector infusion, anti-AAV9 NAbs appeared in blood and CSF. Of note, an IFNγ ELISPOT assay demonstrated the presence of AAV9-specific, but not gigaxonin-specific, Tcells circulating in blood in subjects dosed with the vector and corticosteroids alone (C.G. Bonnemann, 2019, National Center for Advancing Translational Sciences/NIH, conference).

Conclusions

The gene therapy field is experiencing one of its most exciting periods. Long-term efficacy has been achieved in several clinical trials, and gene therapy drug candidates are reaching late-stage clinical development and market approval. Diseases previously untreatable or for which suboptimal treatments were available are now being cured, delivering previously unimaginable outcomes to patients. Since the first reports of immune-related toxicities associated with AAV vector administration, much progress has been made in understanding the interactions between viral vectors and the host immune system. Yet, new unforeseen complexities are still emerging from trials in which high vector doses are administered systemically. This is clear evidence that many questions about AAV vector immunogenicity are still unanswered.

Are All AAV Vectors the Same?

Experience from clinical trials and preclinical models has helped to guide the design of the ideal AAV vector. The best trade-off one can currently imagine is to engineer AAV vectors with better transduction efficiency, carrying optimized therapeutic transgenes and with reduced immunogenic profiles (CpG-depleted genome,66 immunologically inert capsids, vector preparations with low levels of contaminants, minimum amounts of empty capsids226). The ideal vector then would provide a high therapeutic index, as it would permit therapeutic efficacy at doses sufficient to bypass pre-existing humoral immunity, but not above the threshold of activation of Tcell-mediated immunity. While certainly the appetite of the field for novel AAV capsids is still high,227 from an immunological standpoint, with the exception of the potential for NAb escape,228 it is still unclear what are the critical parameters that would make a capsid less likely to trigger immune-mediated toxicities. The fact that AAV vectors used in clinical trials are manufactured with different platforms18 further complicates the analysis of the emerging data.

Are There Any Preclinical Models that Are Useful to Study Anti-AAV Cellular Immune Responses?

Small33,226 and large31 animal models have been useful in studying theimpact of anti-AAV humoral responses and devising strategies around them.119 However, the onset of anti-capsid cellular responses observed in several clinical trials has never been observed before in any of the preclinical animal models employed, even in those susceptible to natural AAV infections such as NHPs,31,229 possibly due to differences in the Tcell compartment when compared to humans.43,230 The lack of relevant animal models remains therefore an important hindrance to fully understanding the biology behind AAV vector immunogenicity.

The first attempts to generate a model via immunization of mice against the AAV capsid were unsuccessful.231, 232, 233 More recently, few mouse models have been established,52,234,235 along with invitro human systems.42,60,64 Although promising, as these models contributed to a better understanding of AAV vector immunogenicity, to date, the efficacy strategies to modulate capsid immune responses can be only tested in the clinic. Animal models, in particular NHPs, remain important to test the safety of any immunomodulatory regimen to be used in combination with AAV gene transfer.91

Are Circulating T Cells Reflective of Immune Responses in Target Organs?

Circulating Tcells are routinely utilized for immunomonitoring in AAV trials (vide infra). However, detection of Tcell reactivity has not always been associated with clinical evidence or immune-mediated toxicities, indicating that observations made in the periphery might not accurately reflect the local immunological events takingplace in the target tissue, where capsid antigens are being presented. While for obvious reasons Tcell responses in peripheral blood remain an important surrogate marker to track vector-related immune responses, analysis of in situ immune responses might help refine our understanding of the mechanisms by which loss or maintenance of transgene expression may occur. Sample collection is likely to be challenging, as clinical manifestations of immune responses in clinical trials have been quite variable in terms of timing, and because often the biomarkers used to monitor these responses are not fully reliable.

What Is The Role of Pre-existing Immunity to WT AAV and Its Interplay with Adaptive Immune Responses to AAV Vectors In Humans?

It is well established that humans carry both humoral and cell-mediated immunity to WT AAV (vide supra). These responses are usually found more frequently in adults, while children are frequently immunologically naive to AAV. Whether the Tcell responses observed in AAV gene transfer trials were primary or memory recall responses has been a matter of debate, as often the delay between vector administration and detection of Tcell activation is incompatible with what is expected for a re-exposure to an antigen encountered in the past. Monitoring immune responses across trials and correlation with the history of exposure to AAV may help in the future to clarify this point. Notably, the recent findings in DMD gene transfer trials, in which complement activation was documented, may be related to the presence of anti-AAV-binding antibodies prior to vector infusion. Thus, in the context of these preliminary observations, it could be informative to carefully assess the history of pre-exposure to WT AAV in relationship to the outcome of gene transfer. Similarly, important information could be gathered by studying what is the effect of activation of innate immune responses triggered by vector administration on memory Tcells that originated from exposure to WT AAV.

Should We Worry About Transgene-Specific Immune Responses and How They Might Impact the Onset of Anti-AAV Immune Responses?

Despite the fact that a wealth of preclinical data are available on immune responses to the transgene product in AAV gene transfer, relatively little information is available on whether this knowledge wouldfaithfully translate to the clinic. Several groups showed that delivery ofAAV vectors to the liver induces transgene-specific tolerance,80, 81, 85, 105 and liver gene transfer has also been used in hemophilia A and B dogs to eradicate anti-F.VIII and F.IX neutralizing antibodies, respectively.76,77 Although these preclinical data on liver tolerance are highly convincing, the open question is whether this concept will reliably translate to humans.

Thus far, no transgene immune responses have been detected in liver trials, although most of the enrolled subjects in these studies were at low risk of counting an anti-transgene immune response. Conversely, transgene-specific immune responses have been detected in trials of CNS gene transfer,88 eye,131 and muscle-directed gene transfer.73,185 Immunosuppression, vector engineering,79 or simply the use of alternate strategies to deliver a therapeutic transgene to a potentially pro-tolerogenic tissue81 are all potential solutions to the issue of transgene immunogenicity. Importantly, the long-term impact of transgene immunity needs to be carefully assessed, particularly on terms of clinical outcomes.86

In clinical trials, anti-transgene immunity will require careful evaluation in patients with pre-existing immunity toward the transgene product, either resulting from the underlying disease73,74 or to previous treatment with recombinant protein.71 Moreover, some diseases involving lysosomal236 or metabolic75 alterations may present with alterations of the immune system, which may interfere with gene transfer.

What Are the Most Promising Approaches to Immunomodulation in AAV Gene Transfer?

Immunomodulation is broadly used in the context of gene transfer to reduce AAV vector immunogenicity, and eventually to allow for repeated interventions. Corticosteroids have been administered in most trials, and they certainly have helped to modulate immune-mediated toxicities and achieve long-term transgene expression. However, in some cases the administration of more complex regimens was required to manage capsid or, potentially, transgene immunogenicity. For vector redosing, several strategies to prevent anti-AAV antibody production have been proposed, with only one tested in the clinic to date (ClinicalTrials.gov: NCT02240407). While some preclinical results are promising, it is clear that pre-existing B cell immunity is a major hurdle, due to the fact that targeting a primed immune system is much harder than simply preventing an immune response to occur.

It is early to identify which strategy or combination of strategies will be the best in the clinic; however, a few general considerations can be made:

  • The ideal immunomodulatory regimen has to offer an acceptable risk-benefit profile, particularly in the context of a given disease indication and patient population (e.g., not all drugs are approved for pediatric use).

  • A careful evaluation of drug interactions should be made to ensure that the co-administration of immunomodulatory drugs does not interfere with the safety and efficacy of gene transfer (e.g., some drugs can interfere with transgene tolerance induction).

  • The ideal immunomodulatory regimen for gene transfer is simple and transient. A short course of clinically approved immunosuppressive drugs may offer an easy path to the clinic and modulate effectively vector-related immune-mediated toxicities.

  • “One size fits all” is unlikely to apply to gene transfer with AAV vectors. Depending on the vector dose, target tissue, disease indication, patient population, and other factors, tailored immunosuppressive regimens will likely have to be identified.

Building Best Practices in Immunomonitoring

As many questions remain on AAV immunogenicity, the field of AAV gene therapy research needs further efforts to resolve the complexity of capsid-related immune responses. The harmonization of patient monitoring using standard guidelines and quality controls, to check immune assay performance over time and across clinical trials, would greatly facilitate the comparison of data, and subsequently the understanding of the complexity of anti-AAV immune responses. Additionally, aside from monitoring IFNγ activation, additional markers of Tcell activation, such as TNF-α,42 which better reflect the profile of activation of Tcells specific to the AAV capsid in humans, should be routinely reported.

The years to come will continue to bring forward a wealth of preclinical and clinical data that hopefully will provide precious insights onsome of the questions we outlined in this review. Gaining a more sophisticated understanding of the AAV vector technology, as well as the nuances of the interactions between AAV vectors and the host immune system, will likely provide a path forward to further extend the success of this still novel and highly promising therapeutic paradigm.

Author Contributions

H.C.V., K.K., and F.M. wrote the manuscript.

Conflicts of Interest

F.M. and K.K. are employees of Spark Therapeutics. H.C.V. declares no competing interests.

Acknowledgments

This work was supported by the European Union’s Research Council (ERC) Consolidator Grant under agreement no. 617432 (MoMAAV to F.M.).

References

1. Lovric J., Mano M., Zentilin L., Eulalio A., Zacchigna S., Giacca M. Terminal differentiation of cardiac and skeletal myocytes induces permissivity to AAV transduction by relieving inhibition imposed by DNA damage response proteins. Mol. Ther. 2012;20:2087–2097. [PMC free article] [PubMed] [Google Scholar]

2. Bennett J., Wellman J., Marshall K.A., McCague S., Ashtari M., DiStefano-Pappas J., Elci O.U., Chung D.C., Sun J., Wright J.F. Safety and durability of effect of contralateral-eye administration of AAV2 gene therapy in patients with childhood-onset blindness caused by RPE65 mutations: a follow-on phase 1 trial. Lancet. 2016;388:661–672. [PMC free article] [PubMed] [Google Scholar]

3. Buchlis G., Podsakoff G.M., Radu A., Hawk S.M., Flake A.W., Mingozzi F., High K.A. Factor IX expression in skeletal muscle of a severe hemophilia B patient 10 years after AAV-mediated gene transfer. Blood. 2012;119:3038–3041. [PMC free article] [PubMed] [Google Scholar]

4. Mueller C., Gernoux G., Gruntman A.M., Borel F., Reeves E.P., Calcedo R., Rouhani F.N., Yachnis A., Humphries M., Campbell-Thompson M. 5 Year expression and neutrophil defect repair after gene therapy in alpha-1 antitrypsin deficiency. Mol. Ther. 2017;25:1387–1394. [PMC free article] [PubMed] [Google Scholar]

5. Nathwani A.C., Reiss U.M., Tuddenham E.G., Rosales C., Chowdary P., McIntosh J., Della Peruta M., Lheriteau E., Patel N., Raj D. Long-term safety and efficacy of factor IX gene therapy in hemophilia B. N.Engl. J. Med. 2014;371:1994–2004. [PMC free article] [PubMed] [Google Scholar]

6. Rossi A., Dupaty L., Aillot L., Zhang L., Gallien C., Hallek M., Odenthal M., Adriouch S., Salvetti A., Büning H. Vector uncoating limits adeno-associated viral vector-mediated transduction of human dendritic cells and vector immunogenicity. Sci. Rep. 2019;9:3631. [PMC free article] [PubMed] [Google Scholar]

7. Somanathan S., Breous E., Bell P., Wilson J.M. AAV vectors avoid inflammatory signals necessary to render transduced hepatocyte targets for destructive Tcells. Mol. Ther. 2010;18:977–982. [PMC free article] [PubMed] [Google Scholar]

8. Gao G., Vandenberghe L.H., Alvira M.R., Lu Y., Calcedo R., Zhou X., Wilson J.M. Clades of adeno-associated viruses are widely disseminated in human tissues. J.Virol. 2004;78:6381–6388. [PMC free article] [PubMed] [Google Scholar]

9. Berns K.I., Muzyczka N. AAV: an overview of unanswered questions. Hum. Gene Ther. 2017;28:308–313. [PMC free article] [PubMed] [Google Scholar]

10. Atchison R.W., Casto B.C., Hammon W.M. Adenovirus-associated defective virus particles. Science. 1965;149:754–756. [PubMed] [Google Scholar]

11. Buller R.M., Janik J.E., Sebring E.D., Rose J.A. Herpes simplex virus types 1 and 2 completely help adenovirus-associated virus replication. J.Virol. 1981;40:241–247. [PMC free article] [PubMed] [Google Scholar]

12. Ogston P., Raj K., Beard P. Productive replication of adeno-associated virus can occur in human papillomavirus type 16 (HPV-16) episome-containing keratinocytes and is augmented by the HPV-16 E2 protein. J.Virol. 2000;74:3494–3504. [PMC free article] [PubMed] [Google Scholar]

13. Moore A.R., Dong B., Chen L., Xiao W. Vaccinia virus as a subhelper for AAV replication and packaging. Mol. Ther. Methods Clin. Dev. 2015;2:15044. [PMC free article] [PubMed] [Google Scholar]

14. Nault J.C., Datta S., Imbeaud S., Franconi A., Mallet M., Couchy G., Letouzé E., Pilati C., Verret B., Blanc J.F. Recurrent AAV2-related insertional mutagenesis in human hepatocellular carcinomas. Nat. Genet. 2015;47:1187–1193. [PubMed] [Google Scholar]

15. La Bella T., Imbeaud S., Peneau C., Mami I., Datta S., Bayard Q., Caruso S., Hirsch T.Z., Calderaro J., Morcrette G. Adeno-associated virus in the liver: natural history and consequences in tumour development. Gut. 2019 Published online August 2, 2019. [PubMed] [Google Scholar]

16. Balakrishnan B., Jayandharan G.R. Basic biology of adeno-associated virus (AAV) vectors used in gene therapy. Curr. Gene Ther. 2014;14:86–100. [PubMed] [Google Scholar]

17. Nonnenmacher M., Weber T. Intracellular transport of recombinant adeno-associated virus vectors. Gene Ther. 2012;19:649–658. [PMC free article] [PubMed] [Google Scholar]

18. Ayuso E., Mingozzi F., Bosch F. Production, purification and characterization of adeno-associated vectors. Curr. Gene Ther. 2010;10:423–436. [PubMed] [Google Scholar]

19. Matsush*ta T., Elliger S., Elliger C., Podsakoff G., Villarreal L., Kurtzman G.J., Iwaki Y., Colosi P. Adeno-associated virus vectors can be efficiently produced without helper virus. Gene Ther. 1998;5:938–945. [PubMed] [Google Scholar]

20. Boutin S., Monteilhet V., Veron P., Leborgne C., Benveniste O., Montus M.F., Masurier C. Prevalence of serum IgG and neutralizing factors against adeno-associated virus (AAV) types 1, 2, 5, 6, 8, and 9 in the healthy population: implications for gene therapy using AAV vectors. Hum. Gene Ther. 2010;21:704–712. [PubMed] [Google Scholar]

21. Calcedo R., Vandenberghe L.H., Gao G., Lin J., Wilson J.M. Worldwide epidemiology of neutralizing antibodies to adeno-associated viruses. J.Infect. Dis. 2009;199:381–390. [PMC free article] [PubMed] [Google Scholar]

22. Calcedo R., Morizono H., Wang L., McCarter R., He J., Jones D., Batshaw M.L., Wilson J.M. Adeno-associated virus antibody profiles in newborns, children, and adolescents. Clin. Vaccine Immunol. 2011;18:1586–1588. [PMC free article] [PubMed] [Google Scholar]

23. Mingozzi F., Chen Y., Edmonson S.C., Zhou S., Thurlings R.M., Tak P.P., High K.A., Vervoordeldonk M.J. Prevalence and pharmacological modulation of humoral immunity to AAV vectors in gene transfer to synovial tissue. Gene Ther. 2013;20:417–424. [PMC free article] [PubMed] [Google Scholar]

24. Perocheau D.P., Cunningham S., Lee J., Antinao Diaz J., Waddington S.N., Gilmour K., Eaglestone S., Lisowski L., Thrasher A.J., Alexander I.E. Age-related seroprevalence of antibodies against AAV-LK03 in a UK population cohort. Hum. Gene Ther. 2019;30:79–87. [PMC free article] [PubMed] [Google Scholar]

25. Fu H., Meadows A.S., Pineda R.J., Kunkler K.L., Truxal K.V., McBride K.L., Flanigan K.M., McCarty D.M. Differential prevalence of antibodies against adeno-associated virus in healthy children and patients with mucopolysaccharidosis III: perspective for AAV-mediated gene therapy. Hum. Gene Ther. Clin. Dev. 2017;28:187–196. [PMC free article] [PubMed] [Google Scholar]

26. Calcedo R., Wilson J.M. AAV natural infection induces broad cross-neutralizing antibody responses to multiple AAV serotypes in chimpanzees. Hum. Gene Ther. Clin. Dev. 2016;27:79–82. [PMC free article] [PubMed] [Google Scholar]

27. Kruzik A., Fetahagic D., Hartlieb B., Dorn S., Koppensteiner H., Horling F.M., Scheiflinger F., Reipert B.M., de la Rosa M. Prevalence of anti-adeno-associated virus immune responses in international cohorts of healthy donors. Mol. Ther. Methods Clin. Dev. 2019;14:126–133. [PMC free article] [PubMed] [Google Scholar]

28. Calcedo R., Wilson J.M. Humoral immune response to AAV. Front. Immunol. 2013;4:341. [PMC free article] [PubMed] [Google Scholar]

29. Leborgne C., Latournerie V., Boutin S., Desgue D., Quéré A., Pignot E., Collaud F., Charles S., Simon Sola M., Masat E. Prevalence and long-term monitoring of humoral immunity against adeno-associated virus in duch*enne muscular dystrophy patients. Cell. Immunol. 2019;342:103780. [PubMed] [Google Scholar]

30. Fitzpatrick Z., Leborgne C., Barbon E., Masat E., Ronzitti G., van Wittenberghe L., Vignaud A., Collaud F., Charles S., Simon Sola M. Influence of pre-existing anti-capsid neutralizing and binding antibodies on AAV vector transduction. Mol. Ther. Methods Clin. Dev. 2018;9:119–129. [PMC free article] [PubMed] [Google Scholar]

31. Jiang H., Couto L.B., Patarroyo-White S., Liu T., Nagy D., Vargas J.A., Zhou S., Scallan C.D., Sommer J., Vijay S. Effects of transient immunosuppression on adenoassociated, virus-mediated, liver-directed gene transfer in rhesus macaques and implications for human gene therapy. Blood. 2006;108:3321–3328. [PMC free article] [PubMed] [Google Scholar]

32. Manno C.S., Pierce G.F., Arruda V.R., Glader B., Ragni M., Rasko J.J., Ozelo M.C., Hoots K., Blatt P., Konkle B. Successful transduction of liver in hemophilia by AAV-factor IX and limitations imposed by the host immune response. Nat. Med. 2006;12:342–347. [PubMed] [Google Scholar]

33. Scallan C.D., Jiang H., Liu T., Patarroyo-White S., Sommer J.M., Zhou S., Couto L.B., Pierce G.F. Human immunoglobulin inhibits liver transduction by AAV vectors at low AAV2 neutralizing titers in SCID mice. Blood. 2006;107:1810–1817. [PubMed] [Google Scholar]

34. Wang L., Calcedo R., Wang H., Bell P., Grant R., Vandenberghe L.H., Sanmiguel J., Morizono H., Batshaw M.L., Wilson J.M. The pleiotropic effects of natural AAV infections on liver-directed gene transfer in macaques. Mol. Ther. 2010;18:126–134. [PMC free article] [PubMed] [Google Scholar]

35. Walport M.J. Complement. First of two parts. N.Engl. J. Med. 2001;344:1058–1066. [PubMed] [Google Scholar]

36. Zaiss A.K., Cotter M.J., White L.R., Clark S.A., Wong N.C., Holers V.M., Bartlett J.S., Muruve D.A. Complement is an essential component of the immune response to adeno-associated virus vectors. J.Virol. 2008;82:2727–2740. [PMC free article] [PubMed] [Google Scholar]

37. Denard J., Marolleau B., Jenny C., Rao T.N., Fehling H.J., Voit T., Svinartchouk F. C-reactive protein (CRP) is essential for efficient systemic transduction of recombinant adeno-associated virus vector 1 (rAAV-1) and rAAV-6 in mice. J.Virol. 2013;87:10784–10791. [PMC free article] [PubMed] [Google Scholar]

38. Murphy S.L., Li H., Mingozzi F., Sabatino D.E., Hui D.J., Edmonson S.A., High K.A. Diverse IgG subclass responses to adeno-associated virus infection and vector administration. J.Med. Virol. 2009;81:65–74. [PMC free article] [PubMed] [Google Scholar]

39. Mingozzi F., Meulenberg J.J., Hui D.J., Basner-Tschakarjan E., Hasbrouck N.C., Edmonson S.A., Hutnick N.A., Betts M.R., Kastelein J.J., Stroes E.S., High K.A. AAV-1-mediated gene transfer to skeletal muscle in humans results in dose-dependent activation of capsid-specific Tcells. Blood. 2009;114:2077–2086. [PMC free article] [PubMed] [Google Scholar]

40. Mingozzi F., Maus M.V., Hui D.J., Sabatino D.E., Murphy S.L., Rasko J.E., Ragni M.V., Manno C.S., Sommer J., Jiang H. CD8+ T-cell responses to adeno-associated virus capsid in humans. Nat. Med. 2007;13:419–422. [PubMed] [Google Scholar]

41. Nathwani A.C., Rosales C., McIntosh J., Rastegarlari G., Nathwani D., Raj D., Nawathe S., Waddington S.N., Bronson R., Jackson S. Long-term safety and efficacy following systemic administration of a self-complementary AAV vector encoding human FIX pseudotyped with serotype 5 and 8 capsid proteins. Mol. Ther. 2011;19:876–885. [PMC free article] [PubMed] [Google Scholar]

42. Kuranda K., Jean-Alphonse P., Leborgne C., Hardet R., Collaud F., Marmier S., Costa Verdera H., Ronzitti G., Veron P., Mingozzi F. Exposure to wild-type AAV drives distinct capsid immunity profiles in humans. J.Clin. Invest. 2018;128:5267–5279. [PMC free article] [PubMed] [Google Scholar]

43. Li H., Lasaro M.O., Jia B., Lin S.W., Haut L.H., High K.A., Ertl H.C. Capsid-specific T-cell responses to natural infections with adeno-associated viruses in humans differ from those of nonhuman primates. Mol. Ther. 2011;19:2021–2030. [PMC free article] [PubMed] [Google Scholar]

44. Xiang Z., Kurupati R.K., Li Y., Kuranda K., Zhou X., Mingozzi F., High K.A., Ertl H.C.J. The effect of CpG sequences on capsid-specific CD8+ Tcell responses to AAV vector gene transfer. Mol. Ther. 2019 Published online November 20, 2019. [PMC free article] [PubMed] [Google Scholar]

45. Hui D.J., Edmonson S.C., Podsakoff G.M., Pien G.C., Ivanciu L., Camire R.M., Ertl H., Mingozzi F., High K.A., Basner-Tschakarjan E. AAV capsid CD8+ T-cell epitopes are highly conserved across AAV serotypes. Mol. Ther. Methods Clin. Dev. 2015;2:15029. [PMC free article] [PubMed] [Google Scholar]

46. Martino A.T., Herzog R.W., Anegon I., Adjali O. Measuring immune responses to recombinant AAV gene transfer. Methods Mol. Biol. 2011;807:259–272. [PMC free article] [PubMed] [Google Scholar]

47. Veron P., Leborgne C., Monteilhet V., Boutin S., Martin S., Moullier P., Masurier C. Humoral and cellular capsid-specific immune responses to adeno-associated virus type 1 in randomized healthy donors. J.Immunol. 2012;188:6418–6424. [PubMed] [Google Scholar]

48. Newell E.W., Sigal N., Bendall S.C., Nolan G.P., Davis M.M. Cytometry by time-of-flight shows combinatorial cytokine expression and virus-specific cell niches within a continuum of CD8+ Tcell phenotypes. Immunity. 2012;36:142–152. [PMC free article] [PubMed] [Google Scholar]

49. Shao W., Earley L.F., Chai Z., Chen X., Sun J., He T., Deng M., Hirsch M.L., Ting J., Samulski R.J., Li C. Double-stranded RNA innate immune response activation from long-term adeno-associated virus vector transduction. JCI Insight. 2018;3:e120474. [PMC free article] [PubMed] [Google Scholar]

50. Trinchieri G., Sher A. Cooperation of Toll-like receptor signals in innate immune defence. Nat. Rev. Immunol. 2007;7:179–190. [PubMed] [Google Scholar]

51. Shirley J.L., Keeler G.D., Sherman A., Zolotukhin I., Markusic D.M., Hoffman B.E., Morel L.M., Wallet, Terhorst C., Herzog R.W. Type I IFN sensing by cDCs and CD4+ Tcell help are both requisite for cross-priming of AAV capsid-specific CD8+ T cells. Mol. Ther. 2019 Published online November 14, 2019. [PMC free article] [PubMed] [Google Scholar]

52. Rogers G.L., Shirley J.L., Zolotukhin I., Kumar S.R.P., Sherman A., Perrin G.Q., Hoffman B.E., Srivastava A., Basner-Tschakarjan E., Wallet M.A. Plasmacytoid and conventional dendritic cells cooperate in crosspriming AAV capsid-specific CD8+ Tcells. Blood. 2017;129:3184–3195. [PMC free article] [PubMed] [Google Scholar]

53. Hösel M., Broxtermann M., Janicki H., Esser K., Arzberger S., Hartmann P., Gillen S., Kleeff J., Stabenow D., Odenthal M. Toll-like receptor 2-mediated innate immune response in human nonparenchymal liver cells toward adeno-associated viral vectors. Hepatology. 2012;55:287–297. [PubMed] [Google Scholar]

54. Martino A.T., Suzuki M., Markusic D.M., Zolotukhin I., Ryals R.C., Moghimi B., Ertl H.C., Muruve D.A., Lee B., Herzog R.W. The genome of self-complementary adeno-associated viral vectors increases Toll-like receptor 9-dependent innate immune responses in the liver. Blood. 2011;117:6459–6468. [PMC free article] [PubMed] [Google Scholar]

55. Zhu J., Huang X., Yang Y. The TLR9-MyD88 pathway is critical for adaptive immune responses to adeno-associated virus gene therapy vectors in mice. J.Clin. Invest. 2009;119:2388–2398. [PMC free article] [PubMed] [Google Scholar]

56. Herzog R.W., Cooper M., Perrin G.Q., Biswas M., Martino A.T., Morel L., Terhorst C., Hoffman B.E. Regulatory Tcells and TLR9 activation shape antibody formation to a secreted transgene product in AAV muscle gene transfer. Cell. Immunol. 2019;342:103682. [PMC free article] [PubMed] [Google Scholar]

57. Rogers G.L., Suzuki M., Zolotukhin I., Markusic D.M., Morel L.M., Lee B., Ertl H.C., Herzog R.W. Unique roles of TLR9- and MyD88-dependent and -independent pathways in adaptive immune responses to AAV-mediated gene transfer. J.Innate Immun. 2015;7:302–314. [PMC free article] [PubMed] [Google Scholar]

58. Ashley S.N., Somanathan S., Giles A.R., Wilson J.M. TLR9 signaling mediates adaptive immunity following systemic AAV gene therapy. Cell. Immunol. 2019;346:103997. [PMC free article] [PubMed] [Google Scholar]

59. Abbas A.K., Lichtman A.H. Saunders/Elsevier; 2009. Basic Immunology: Functions and Disorders of the Immune System. [Google Scholar]

60. Pien G.C., Basner-Tschakarjan E., Hui D.J., Mentlik A.N., Finn J.D., Hasbrouck N.C., Zhou S., Murphy S.L., Maus M.V., Mingozzi F. Capsid antigen presentation flags human hepatocytes for destruction after transduction by adeno-associated viral vectors. J.Clin. Invest. 2009;119:1688–1695. [PMC free article] [PubMed] [Google Scholar]

61. Palaschak B., Marsic D., Herzog R.W., Zolotukhin S., Markusic D.M. An immune-competent murine model to study elimination of AAV-transduced hepatocytes by capsid-specific CD8+ Tcells. Mol. Ther. Methods Clin. Dev. 2017;5:142–152. [PMC free article] [PubMed] [Google Scholar]

62. George L.A., Sullivan S.K., Giermasz A., Rasko J.E.J., Samelson-Jones B.J., Ducore J., Cuker A., Sullivan L.M., Majumdar S., Teitel J. Hemophilia B gene therapy with a high-specific-activity factor IX variant. N.Engl. J. Med. 2017;377:2215–2227. [PMC free article] [PubMed] [Google Scholar]

63. Nathwani A.C., Tuddenham E.G., Rangarajan S., Rosales C., McIntosh J., Linch D.C., Chowdary P., Riddell A., Pie A.J., Harrington C. Adenovirus-associated virus vector-mediated gene transfer in hemophilia B. N.Engl. J. Med. 2011;365:2357–2365. [PMC free article] [PubMed] [Google Scholar]

64. Finn J.D., Hui D., Downey H.D., Dunn D., Pien G.C., Mingozzi F., Zhou S., High K.A. Proteasome inhibitors decrease AAV2 capsid derived peptide epitope presentation on MHC class I following transduction. Mol. Ther. 2010;18:135–142. [PMC free article] [PubMed] [Google Scholar]

65. Matzinger P. The danger model: a renewed sense of self. Science. 2002;296:301–305. [PubMed] [Google Scholar]

66. Faust S.M., Bell P., Cutler B.J., Ashley S.N., Zhu Y., Rabinowitz J.E., Wilson J.M. CpG-depleted adeno-associated virus vectors evade immune detection. J.Clin. Invest. 2013;123:2994–3001. [PMC free article] [PubMed] [Google Scholar]

67. Butterfield J.S.S., Biswas M., Shirley J.L., Kumar S.R.P., Sherman A., Terhorst C., Ling C., Herzog R.W. TLR9-activating CpG-B ODN but not TLR7 agonists triggers antibody formation to factor IX in muscle gene transfer. Hum. Gene Ther. Methods. 2019;30:81–92. [PMC free article] [PubMed] [Google Scholar]

68. Hoffman B.E., Martino A.T., Sack B.K., Cao O., Liao G., Terhorst C., Herzog R.W. Nonredundant roles of IL-10 and TGF-β in suppression of immune responses to hepatic AAV-factor IX gene transfer. Mol. Ther. 2011;19:1263–1272. [PMC free article] [PubMed] [Google Scholar]

69. Poupiot J., Costa Verdera H., Hardet R., Colella P., Collaud F., Bartolo L., Davoust J., Sanatine P., Mingozzi F., Richard I., Ronzitti G. Role of regulatory Tcell and effector Tcell exhaustion in liver-mediated transgene tolerance in muscle. Mol. Ther. Methods Clin. Dev. 2019;15:83–100. [PMC free article] [PubMed] [Google Scholar]

70. Xiong W., Wu D.M., Xue Y., Wang S.K., Chung M.J., Ji X., Rana P., Zhao S.R., Mai S., Cepko C.L. AAV cis-regulatory sequences are correlated with ocular toxicity. Proc. Natl. Acad. Sci. USA. 2019;116:5785–5794. [PMC free article] [PubMed] [Google Scholar]

71. Kishnani P.S., Goldenberg P.C., DeArmey S.L., Heller J., Benjamin D., Young S., Bali D., Smith S.A., Li J.S., Mandel H. Cross-reactive immunologic material status affects treatment outcomes in Pompe disease infants. Mol. Genet. Metab. 2010;99:26–33. [PMC free article] [PubMed] [Google Scholar]

72. Fields P.A., Arruda V.R., Armstrong E., Chu K., Mingozzi F., Hagstrom J.N., Herzog R.W., High K.A. Risk and prevention of anti-factor IX formation in AAV-mediated gene transfer in the context of a large deletion of F9. Mol. Ther. 2001;4:201–210. [PubMed] [Google Scholar]

73. Mendell J.R., Campbell K., Rodino-Klapac L., Sahenk Z., Shilling C., Lewis S., Bowles D., Gray S., Li C., Galloway G. Dystrophin immunity in duch*enne’s muscular dystrophy. N.Engl. J. Med. 2010;363:1429–1437. [PMC free article] [PubMed] [Google Scholar]

74. Flanigan K.M., Campbell K., Viollet L., Wang W., Gomez A.M., Walker C.M., Mendell J.R. Anti-dystrophin Tcell responses in duch*enne muscular dystrophy: prevalence and a glucocorticoid treatment effect. Hum. Gene Ther. 2013;24:797–806. [PMC free article] [PubMed] [Google Scholar]

75. Melis D., Carbone F., Minopoli G., La Rocca C., Perna F., De Rosa V., Galgani M., Andria G., Parenti G., Matarese G. Cutting edge: increased autoimmunity risk in glycogen storage disease type 1b is associated with a reduced engagement of glycolysis in Tcells and an impaired regulatory Tcell function. J.Immunol. 2017;198:3803–3808. [PMC free article] [PubMed] [Google Scholar]

76. Crudele J.M., Finn J.D., Siner J.I., Martin N.B., Niemeyer G.P., Zhou S., Mingozzi F., Lothrop C.D., Jr., Arruda V.R. AAV liver expression of FIX-Padua prevents and eradicates FIX inhibitor without increasing thrombogenicity in hemophilia B dogs and mice. Blood. 2015;125:1553–1561. [PMC free article] [PubMed] [Google Scholar]

77. Finn J.D., Ozelo M.C., Sabatino D.E., Franck H.W., Merricks E.P., Crudele J.M., Zhou S., Kazazian H.H., Lillicrap D., Nichols T.C., Arruda V.R. Eradication of neutralizing antibodies to factor VIII in canine hemophilia A after liver gene therapy. Blood. 2010;116:5842–5848. [PMC free article] [PubMed] [Google Scholar]

78. Herzog R.W., Fields P.A., Arruda V.R., Brubaker J.O., Armstrong E., McClintock D., Bellinger D.A., Couto L.B., Nichols T.C., High K.A. Influence of vector dose on factor IX-specific T and B cell responses in muscle-directed gene therapy. Hum. Gene Ther. 2002;13:1281–1291. [PubMed] [Google Scholar]

79. Colella P., Sellier P., Costa Verdera H., Puzzo F., van Wittenberghe L., Guerchet N., Daniele N., Gjata B., Marmier S., Charles S. AAV gene transfer with tandem promoter design prevents anti-transgene immunity and provides persistent efficacy in neonate Pompe mice. Mol. Ther. Methods Clin. Dev. 2018;12:85–101. [PMC free article] [PubMed] [Google Scholar]

80. Mingozzi F., Liu Y.L., Dobrzynski E., Kaufhold A., Liu J.H., Wang Y., Arruda V.R., High K.A., Herzog R.W. Induction of immune tolerance to coagulation factor IX antigen by invivo hepatic gene transfer. J.Clin. Invest. 2003;111:1347–1356. [PMC free article] [PubMed] [Google Scholar]

81. Puzzo F., Colella P., Biferi M.G., Bali D., Paulk N.K., Vidal P., Collaud F., Simon-Sola M., Charles S., Hardet R. Rescue of Pompe disease in mice by AAV-mediated liver delivery of secretable acid α-glucosidase. Sci. Transl. Med. 2017;9:eaam6375. [PMC free article] [PubMed] [Google Scholar]

82. Perrin G.Q., Zolotukhin I., Sherman A., Biswas M., de Jong Y.P., Terhorst C., Davidoff A.M., Herzog R.W. Dynamics of antigen presentation to transgene product-specific CD4+ Tcells and of Treg induction upon hepatic AAV gene transfer. Mol. Ther. Methods Clin. Dev. 2016;3:16083. [PMC free article] [PubMed] [Google Scholar]

83. Dasgupta S., Navarrete A.M., Bayry J., Delignat S., Wootla B., André S., Christophe O., Nascimbeni M., Jacquemin M., Martinez-Pomares L. A role for exposed mannosylations in presentation of human therapeutic self-proteins to CD4+ T lymphocytes. Proc. Natl. Acad. Sci. USA. 2007;104:8965–8970. [PMC free article] [PubMed] [Google Scholar]

84. Jirovska D., Ye P., Pipe S.W., Miao C.H. Reduction of inhibitory anti-FVIII antibody titer by using a B domain variant FVIII/N6 cDNA for nonviral gene therapy in hemophilia A mice. Blood. 2008;112:3537. [Google Scholar]

85. Kumar S.R.P., Hoffman B.E., Terhorst C., de Jong Y.P., Herzog R.W. The balance between CD8+ Tcell-mediated clearance of AAV-encoded antigen in the liver and tolerance is dependent on the vector dose. Mol. Ther. 2017;25:880–891. [PMC free article] [PubMed] [Google Scholar]

86. Calcedo R., Somanathan S., Qin Q., Betts M.R., Rech A.J., Vonderheide R.H., Mueller C., Flotte T.R., Wilson J.M. Class I-restricted T-cell responses to a polymorphic peptide in a gene therapy clinical trial for α-1-antitrypsin deficiency. Proc. Natl. Acad. Sci. USA. 2017;114:1655–1659. [PMC free article] [PubMed] [Google Scholar]

87. Servais L., Shieh P., Dowling J., Kuntz N., Müller-Felber W., Smith B. P.105INCEPTUS pre-phase 1, prospective, non-interventional, natural history run-in study to evaluate subjects aged 4 years and younger with X-linked myotubular myopathy (XLMTM) Neuromuscul. Disord. 2019;29:S79. [Google Scholar]

88. Tardieu M., Zérah M., Gougeon M.L., Ausseil J., de Bournonville S., Husson B., Zafeiriou D., Parenti G., Bourget P., Poirier B. Intracerebral gene therapy in children with mucopolysaccharidosis type IIIB syndrome: an uncontrolled phase 1/2 clinical trial. Lancet Neurol. 2017;16:712–720. [PubMed] [Google Scholar]

89. Boisgerault F., Mingozzi F. The skeletal muscle environment and its role in immunity and tolerance to AAV vector-mediated gene transfer. Curr. Gene Ther. 2015;15:381–394. [PMC free article] [PubMed] [Google Scholar]

90. Cao O., Dobrzynski E., Wang L., Nayak S., Mingle B., Terhorst C., Herzog R.W. Induction and role of regulatory CD4+CD25+ Tcells in tolerance to the transgene product following hepatic invivo gene transfer. Blood. 2007;110:1132–1140. [PMC free article] [PubMed] [Google Scholar]

91. Mingozzi F., Hasbrouck N.C., Basner-Tschakarjan E., Edmonson S.A., Hui D.J., Sabatino D.E., Zhou S., Wright J.F., Jiang H., Pierce G.F. Modulation of tolerance to the transgene product in a nonhuman primate model of AAV-mediated gene transfer to liver. Blood. 2007;110:2334–2341. [PMC free article] [PubMed] [Google Scholar]

92. Sherman A., Biswas M., Herzog R.W. Innovative Approaches for Immune Tolerance to Factor VIII in the Treatment of Hemophilia A. Front Immunol. 2017;8:1604. [PMC free article] [PubMed] [Google Scholar]

93. Pillarisetty V.G., Shah A.B., Miller G., Bleier J.I., DeMatteo R.P. Liver dendritic cells are less immunogenic than spleen dendritic cells because of differences in subtype composition. J.Immunol. 2004;172:1009–1017. [PubMed] [Google Scholar]

94. You Q., Cheng L., Kedl R.M., Ju C. Mechanism of Tcell tolerance induction by murine hepatic Kupffer cells. Hepatology. 2008;48:978–990. [PMC free article] [PubMed] [Google Scholar]

95. Katz S.C., Pillarisetty V.G., Bleier J.I., Shah A.B., DeMatteo R.P. Liver sinusoidal endothelial cells are insufficient to activate Tcells. J.Immunol. 2004;173:230–235. [PubMed] [Google Scholar]

96. Knolle P., Schlaak J., Uhrig A., Kempf P., Meyer zum Büschenfelde K.H., Gerken G. Human Kupffer cells secrete IL-10 in response to lipopolysaccharide (LPS) challenge. J.Hepatol. 1995;22:226–229. [PubMed] [Google Scholar]

97. Liu J., Yu Q., Wu W., Huang X., Broering R., Werner M., Roggendorf M., Yang D., Lu M. TLR2 stimulation strengthens intrahepatic myeloid-derived cell-mediated T Cell tolerance through inducing Kupffer cell expansion and IL-10 production. J.Immunol. 2018;200:2341–2351. [PubMed] [Google Scholar]

98. Breous E., Somanathan S., Vandenberghe L.H., Wilson J.M. Hepatic regulatory Tcells and Kupffer cells are crucial mediators of systemic Tcell tolerance to antigens targeting murine liver. Hepatology. 2009;50:612–621. [PMC free article] [PubMed] [Google Scholar]

99. Burghardt S., Claass B., Erhardt A., Karimi K., Tiegs G. Hepatocytes induce Foxp3+ regulatory Tcells by Notch signaling. J.Leukoc. Biol. 2014;96:571–577. [PubMed] [Google Scholar]

100. Burghardt S., Erhardt A., Claass B., Huber S., Adler G., Jacobs T., Chalaris A., Schmidt-Arras D., Rose-John S., Karimi K., Tiegs G. Hepatocytes contribute to immune regulation in the liver by activation of the Notch signaling pathway in Tcells. J.Immunol. 2013;191:5574–5582. [PubMed] [Google Scholar]

101. Breous E., Somanathan S., Wilson J.M. BALB/c mice show impaired hepatic tolerogenic response following AAV gene transfer to the liver. Mol. Ther. 2010;18:766–774. [PMC free article] [PubMed] [Google Scholar]

102. Battaglia M., Stabilini A., Roncarolo M.G. Rapamycin selectively expands CD4+CD25+FoxP3+ regulatory Tcells. Blood. 2005;105:4743–4748. [PubMed] [Google Scholar]

103. Biswas M., Sarkar D., Kumar S.R.P., Nayak S., Rogers G.L., Markusic D.M., Liao G., Terhorst C., Herzog R.W. Synergy between rapamycin and FLT3 ligand enhances plasmacytoid dendritic cell-dependent induction of CD4+CD25+FoxP3+ Treg. Blood. 2015;125:2937–2947. [PMC free article] [PubMed] [Google Scholar]

104. Keeler G.D., Kumar S., Palaschak B., Silverberg E.L., Markusic D.M., Jones N.T., Hoffman B.E. Gene therapy-induced antigen-specific Tregs inhibit neuro-inflammation and reverse disease in a mouse model of multiple sclerosis. Mol. Ther. 2018;26:173–183. [PMC free article] [PubMed] [Google Scholar]

105. Dobrzynski E., Mingozzi F., Liu Y.-L., Bendo E., Cao O., Wang L., Herzog R.W. Induction of antigen-specific CD4+ T-cell anergy and deletion by invivo viral gene transfer. Blood. 2004;104:969–977. [PubMed] [Google Scholar]

106. Holz L.E., Benseler V., Bowen D.G., Bouillet P., Strasser A., O’Reilly L., d’Avigdor W.M., Bishop A.G., McCaughan G.W., Bertolino P. Intrahepatic murine CD8 T-cell activation associates with a distinct phenotype leading to Bim-dependent death. Gastroenterology. 2008;135:989–997. [PMC free article] [PubMed] [Google Scholar]

107. Benseler V., Warren A., Vo M., Holz L.E., Tay S.S., Le Couteur D.G., Breen E., Allison A.C., van Rooijen N., McGuffog C. Hepatocyte entry leads to degradation of autoreactive CD8 Tcells. Proc. Natl. Acad. Sci. USA. 2011;108:16735–16740. [PMC free article] [PubMed] [Google Scholar]

108. Le Guen V., Judor J.P., Boeffard F., Gauttier V., Ferry N., Soulillou J.P., Brouard S., Conchon S. Alloantigen gene transfer to hepatocytes promotes tolerance to pancreatic islet graft by inducing CD8+ regulatory Tcells. J.Hepatol. 2017;66:765–777. [PubMed] [Google Scholar]

109. Bartolo L., Li Chung Tong S., Chappert P., Urbain D., Collaud F., Colella P., Richard I., Ronzitti G., Demengeot J., Gross D.A. Dual muscle-liver transduction imposes immune tolerance for muscle transgene engraftment despite preexisting immunity. JCI Insight. 2019;4:4. [PMC free article] [PubMed] [Google Scholar]

110. Bénéchet A.P., De Simone G., Di Lucia P., Cilenti F., Barbiera G., Le Bert N., Fumagalli V., Lusito E., Moalli F., Bianchessi V. Dynamics and genomic landscape of CD8+ Tcells undergoing hepatic priming. Nature. 2019;574:200–205. [PMC free article] [PubMed] [Google Scholar]

111. Markusic D.M., Hoffman B.E., Perrin G.Q., Nayak S., Wang X., LoDuca P.A., High K.A., Herzog R.W. Effective gene therapy for haemophilic mice with pathogenic factor IX antibodies. EMBO Mol. Med. 2013;5:1698–1709. [PMC free article] [PubMed] [Google Scholar]

112. Tay S.S., Wong Y.C., McDonald D.M., Wood N.A., Roediger B., Sierro F., Mcguffog C., Alexander I.E., Bishop G.A., Gamble J.R. Antigen expression level threshold tunes the fate of CD8 Tcells during primary hepatic immune responses. Proc. Natl. Acad. Sci. USA. 2014;111:E2540–E2549. [PMC free article] [PubMed] [Google Scholar]

113. Han S.O., Ronzitti G., Arnson B., Leborgne C., Li S., Mingozzi F., Koeberl D. Low-dose liver-targeted gene therapy for Pompe disease enhances therapeutic efficacy of ERT via immune tolerance induction. Mol. Ther. Methods Clin. Dev. 2017;4:126–136. [PMC free article] [PubMed] [Google Scholar]

114. Masat E., Pavani G., Mingozzi F. Humoral immunity to AAV vectors in gene therapy: challenges and potential solutions. Discov. Med. 2013;15:379–389. [PubMed] [Google Scholar]

115. Colella P., Ronzitti G., Mingozzi F. Emerging Issues in AAV-mediated invivo gene therapy. Mol. Ther. Methods Clin. Dev. 2017;8:87–104. [PMC free article] [PubMed] [Google Scholar]

116. Corti M., Cleaver B., Clément N., Conlon T.J., Faris K.J., Wang G., Benson J., Tarantal A.F., Fuller D., Herzog R.W., Byrne B.J. Evaluation of readministration of a recombinant adeno-associated virus vector expressing acid alpha-glucosidase in Pompe disease: preclinical to clinical planning. Hum. Gene Ther. Clin. Dev. 2015;26:185–193. [PMC free article] [PubMed] [Google Scholar]

117. Wang L., Wang H., Bell P., McMenamin D., Wilson J.M. Hepatic gene transfer in neonatal mice by adeno-associated virus serotype 8 vector. Hum. Gene Ther. 2012;23:533–539. [PMC free article] [PubMed] [Google Scholar]

118. Manno C.S., Chew A.J., Hutchison S., Larson P.J., Herzog R.W., Arruda V.R., Tai S.J., Ragni M.V., Thompson A., Ozelo M. AAV-mediated factor IX gene transfer to skeletal muscle in patients with severe hemophilia B. Blood. 2003;101:2963–2972. [PubMed] [Google Scholar]

119. Meliani A., Boisgerault F., Hardet R., Marmier S., Collaud F., Ronzitti G., Leborgne C., Costa Verdera H., Simon Sola M., Charles S. Antigen-selective modulation of AAV immunogenicity with tolerogenic rapamycin nanoparticles enables successful vector re-administration. Nat. Commun. 2018;9:4098. [PMC free article] [PubMed] [Google Scholar]

120. Streilein J.W. Ocular immune privilege: therapeutic opportunities from an experiment of nature. Nat. Rev. Immunol. 2003;3:879–889. [PubMed] [Google Scholar]

121. Spadoni I., Fornasa G., Rescigno M. Organ-specific protection mediated by cooperation between vascular and epithelial barriers. Nat. Rev. Immunol. 2017;17:761–773. [PubMed] [Google Scholar]

122. Martínez-Alcantar L., Talavera-Carrillo D.K., Pineda-Salazar J.U., Ávalos-Viveros M., Gutiérrez-Ospina G., Phillips-Farfán B.V., Fuentes-Farías A.L., Meléndez-Herrera E. Anterior chamber associated immune deviation to cytosolic neural antigens avoids self-reactivity after optic nerve injury and polarizes the retinal environment to an anti-inflammatory profile. J.Neuroimmunol. 2019;333:476964. [PubMed] [Google Scholar]

123. Keino H., Horie S., Sugita S. Immune privilege and eye-derived T-regulatory cells. J.Immunol. Res. 2018;2018:1679197. [PMC free article] [PubMed] [Google Scholar]

124. Anand V., Duffy B., Yang Z., Dejneka N.S., Maguire A.M., Bennett J. A deviant immune response to viral proteins and transgene product is generated on subretinal administration of adenovirus and adeno-associated virus. Mol. Ther. 2002;5:125–132. [PubMed] [Google Scholar]

125. Amado D., Mingozzi F., Hui D., Bennicelli J.L., Wei Z., Chen Y., Bote E., Grant R.L., Golden J.A., Narfstrom K. Safety and efficacy of subretinal readministration of a viral vector in large animals to treat congenital blindness. Sci. Transl. Med. 2010;2:21ra16. [PMC free article] [PubMed] [Google Scholar]

126. Bainbridge J.W., Smith A.J., Barker S.S., Robbie S., Henderson R., Balaggan K., Viswanathan A., Holder G.E., Stockman A., Tyler N. Effect of gene therapy on visual function in Leber’s congenital amaurosis. N.Engl. J. Med. 2008;358:2231–2239. [PubMed] [Google Scholar]

127. Maguire A.M., Simonelli F., Pierce E.A., Pugh E.N., Jr., Mingozzi F., Bennicelli J., Banfi S., Marshall K.A., Testa F., Surace E.M. Safety and efficacy of gene transfer for Leber’s congenital amaurosis. N.Engl. J. Med. 2008;358:2240–2248. [PMC free article] [PubMed] [Google Scholar]

128. Bennett J. Taking stock of retinal gene therapy: looking back and moving forward. Mol. Ther. 2017;25:1076–1094. [PMC free article] [PubMed] [Google Scholar]

129. Li Q., Miller R., Han P.Y., Pang J., Dinculescu A., Chiodo V., Hauswirth W.W. Intraocular route of AAV2 vector administration defines humoral immune response and therapeutic potential. Mol. Vis. 2008;14:1760–1769. [PMC free article] [PubMed] [Google Scholar]

130. Hauswirth W.W., Aleman T.S., Kaushal S., Cideciyan A.V., Schwartz S.B., Wang L., Conlon T.J., Boye S.L., Flotte T.R., Byrne B.J., Jacobson S.G. Treatment of Leber congenital amaurosis due to RPE65 mutations by ocular subretinal injection of adeno-associated virus gene vector: short-term results of a phase I trial. Hum. Gene Ther. 2008;19:979–990. [PMC free article] [PubMed] [Google Scholar]

131. Le Meur G., Lebranchu P., Billaud F., Adjali O., Schmitt S., Bézieau S., Péréon Y., Valabregue R., Ivan C., Darmon C. Safety and long-term efficacy of AAV4 gene therapy in patients with RPE65 Leber congenital amaurosis. Mol. Ther. 2018;26:256–268. [PMC free article] [PubMed] [Google Scholar]

132. Bainbridge J.W., Mehat M.S., Sundaram V., Robbie S.J., Barker S.E., Ripamonti C., Georgiadis A., Mowat F.M., Beattie S.G., Gardner P.J. Long-term effect of gene therapy on Leber’s congenital amaurosis. N.Engl. J. Med. 2015;372:1887–1897. [PMC free article] [PubMed] [Google Scholar]

133. Jacobson S.G., Cideciyan A.V., Ratnakaram R., Heon E., Schwartz S.B., Roman A.J., Peden M.C., Aleman T.S., Boye S.L., Sumaroka A. Gene therapy for Leber congenital amaurosis caused by RPE65 mutations: safety and efficacy in 15 children and adults followed up to 3 years. Arch. Ophthalmol. 2012;130:9–24. [PMC free article] [PubMed] [Google Scholar]

134. Ashtari M., Cyckowski L.L., Monroe J.F., Marshall K.A., Chung D.C., Auricchio A., Simonelli F., Leroy B.P., Maguire A.M., Shindler K.S., Bennett J. The human visual cortex responds to gene therapy-mediated recovery of retinal function. J.Clin. Invest. 2011;121:2160–2168. [PMC free article] [PubMed] [Google Scholar]

135. Maguire A.M., High K.A., Auricchio A., Wright J.F., Pierce E.A., Testa F., Mingozzi F., Bennicelli J.L., Ying G.S., Rossi S. Age-dependent effects of RPE65 gene therapy for Leber’s congenital amaurosis: a phase 1 dose-escalation trial. Lancet. 2009;374:1597–1605. [PMC free article] [PubMed] [Google Scholar]

136. Bennett J., Ashtari M., Wellman J., Marshall K.A., Cyckowski L.L., Chung D.C., McCague S., Pierce E.A., Chen Y., Bennicelli J.L. AAV2 gene therapy readministration in three adults with congenital blindness. Sci. Transl. Med. 2012;4 120ra115. [PMC free article] [PubMed] [Google Scholar]

137. Maguire A.M., Russell S., Wellman J.A., Chung D.C., Yu Z.F., Tillman A., Wittes J., Pappas J., Elci O., Marshall K.A. Efficacy, safety, and durability of voretigene neparvovec-rzyl in RPE65 mutation-associated inherited retinal dystrophy: results of phase 1 and 3 trials. Ophthalmology. 2019;126:1273–1285. [PubMed] [Google Scholar]

138. Russell S., Bennett J., Wellman J.A., Chung D.C., Yu Z.F., Tillman A., Wittes J., Pappas J., Elci O., McCague S. Efficacy and safety of voretigene neparvovec (AAV2-hRPE65v2) in patients with RPE65-mediated inherited retinal dystrophy: a randomised, controlled, open-label, phase 3 trial. Lancet. 2017;390:849–860. [PMC free article] [PubMed] [Google Scholar]

139. Weleber R.G., Pennesi M.E., Wilson D.J., Kaushal S., Erker L.R., Jensen L., McBride M.T., Flotte T.R., Humphries M., Calcedo R. Results at 2 years after gene therapy for RPE65-deficient Leber congenital amaurosis and severe early-childhood-onset retinal dystrophy. Ophthalmology. 2016;123:1606–1620. [PubMed] [Google Scholar]

140. Pennesi M.E., Weleber R.G., Yang P., Whitebirch C., Thean B., Flotte T.R., Humphries M., Chegarnov E., Beasley K.N., Stout J.T., Chulay J.D. Results at 5 years after gene therapy for RPE65-deficient retinal dystrophy. Hum. Gene Ther. 2018;29:1428–1437. [PubMed] [Google Scholar]

141. Jacobson S.G., Acland G.M., Aguirre G.D., Aleman T.S., Schwartz S.B., Cideciyan A.V., Zeiss C.J., Komaromy A.M., Kaushal S., Roman A.J. Safety of recombinant adeno-associated virus type 2-RPE65 vector delivered by ocular subretinal injection. Mol. Ther. 2006;13:1074–1084. [PubMed] [Google Scholar]

142. Jacobson S.G., Cideciyan A.V., Roman A.J., Sumaroka A., Schwartz S.B., Heon E., Hauswirth W.W. Improvement and decline in vision with gene therapy in childhood blindness. N.Engl. J. Med. 2015;372:1920–1926. [PMC free article] [PubMed] [Google Scholar]

143. Testa F., Maguire A.M., Rossi S., Pierce E.A., Melillo P., Marshall K., Banfi S., Surace E.M., Sun J., Acerra C. Three-year follow-up after unilateral subretinal delivery of adeno-associated virus in patients with Leber congenital Amaurosis type 2. Ophthalmology. 2013;120:1283–1291. [PMC free article] [PubMed] [Google Scholar]

144. Darrow J.J. Luxturna: FDA documents reveal the value of a costly gene therapy. Drug Discov. Today. 2019;24:949–954. [PubMed] [Google Scholar]

145. Trapani I., Auricchio A. Has retinal gene therapy come of age? From bench to bedside and back to bench. Hum. Mol. Genet. 2019;28(R1):R108–R118. [PMC free article] [PubMed] [Google Scholar]

146. Fischer M.D., Ochakovski G.A., Beier B., Seitz I.P., Vaheb Y., Kortuem C., Reichel F.F.L., Kuehlewein L., Kahle N.A., Peters T. Efficacy and safety of retinal gene therapy using adeno-associated virus vector for patients with choroideremia: a randomized clinical trial. JAMA Ophthalmol. 2019;137:1247–1254. [PMC free article] [PubMed] [Google Scholar]

147. MacLaren R.E., Groppe M., Barnard A.R., Cottriall C.L., Tolmachova T., Seymour L., Clark K.R., During M.J., Cremers F.P., Black G.C. Retinal gene therapy in patients with choroideremia: initial findings from a phase 1/2 clinical trial. Lancet. 2014;383:1129–1137. [PMC free article] [PubMed] [Google Scholar]

148. Ghazi N.G., Abboud E.B., Nowilaty S.R., Alkuraya H., Alhommadi A., Cai H., Hou R., Deng W.T., Boye S.L., Almaghamsi A. Treatment of retinitis pigmentosa due to MERTK mutations by ocular subretinal injection of adeno-associated virus gene vector: results of a phase I trial. Hum. Genet. 2016;135:327–343. [PubMed] [Google Scholar]

149. Rakoczy E.P., Lai C.M., Magno A.L., Wikstrom M.E., French M.A., Pierce C.M., Schwartz S.D., Blumenkranz M.S., Chalberg T.W., Degli-Esposti M.A., Constable I.J. Gene therapy with recombinant adeno-associated vectors for neovascular age-related macular degeneration: 1 year follow-up of a phase 1 randomised clinical trial. Lancet. 2015;386:2395–2403. [PubMed] [Google Scholar]

150. Constable I.J., Blumenkranz M.S., Schwartz S.D., Barone S., Lai C.M., Rakoczy E.P. Gene therapy for age-related macular degeneration. Asia Pac. J. Ophthalmol. (Phila.) 2016;5:300–303. [PubMed] [Google Scholar]

151. Constable I.J., Pierce C.M., Lai C.M., Magno A.L., Degli-Esposti M.A., French M.A., McAllister I.L., Butler S., Barone S.B., Schwartz S.D. Phase 2a randomized clinical trial: safety and post hoc analysis of subretinal rAAV.sFLT-1 for wet age-related macular degeneration. EBioMedicine. 2016;14:168–175. [PMC free article] [PubMed] [Google Scholar]

152. Chan Y.K., Wang S., Letizia A., Chan Y., Lim E., Graveline A., Costa Verdera H., Alphonse P., Xue Y., Chiang J. Reducing AAV-mediated immune responses and pathology in a subretinal pig model by engineering the vector genome. Mol. Ther. 2019;27:298. [Google Scholar]

153. Trapani I., Colella P., Sommella A., Iodice C., Cesi G., de Simone S., Marrocco E., Rossi S., Giunti M., Palfi A. Effective delivery of large genes to the retina by dual AAV vectors. EMBO Mol. Med. 2014;6:194–211. [PMC free article] [PubMed] [Google Scholar]

154. Hung S.S.C., Chrysostomou V., Li F., Lim J.K.H., Wang J.-H., Powell J.E., Tu L., Daniszewski M., Lo C., Wong R.C. AAV-mediated CRISPR/Cas gene editing of retinal cells invivo. Invest. Ophthalmol. Vis. Sci. 2016;57:3470–3476. [PubMed] [Google Scholar]

155. Moreno A.M., Fu X., Zhu J., Katrekar D., Shih Y.V., Marlett J., Cabotaje J., Tat J., Naughton J., Lisowski L. In situ gene therapy via AAV-CRISPR-Cas9-mediated targeted gene regulation. Mol. Ther. 2018;26:1818–1827. [PMC free article] [PubMed] [Google Scholar]

156. Yu W., Mookherjee S., Chaitankar V., Hiriyanna S., Kim J.-W., Brooks M., Ataeijannati Y., Sun X., Dong L., Li T. Nrl knockdown by AAV-delivered CRISPR/Cas9 prevents retinal degeneration in mice. Nat. Commun. 2017;8:14716. [PMC free article] [PubMed] [Google Scholar]

157. Jo D.H., Koo T., Cho C.S., Kim J.H., Kim J.S., Kim J.H. Long-term effects of invivo genome editing in the mouse retina using Campylobacter jejuni Cas9 expressed via adeno-associated virus. Mol. Ther. 2019;27:130–136. [PMC free article] [PubMed] [Google Scholar]

158. Crudele J.M., Chamberlain J.S. Cas9 immunity creates challenges for CRISPR gene editing therapies. Nat. Commun. 2018;9:3497. [PMC free article] [PubMed] [Google Scholar]

159. Wagner D.L., Amini L., Wendering D.J., Burkhardt L.M., Akyüz L., Reinke P., Volk H.D., Schmueck-Henneresse M. High prevalence of Streptococcus pyogenes Cas9-reactive Tcells within the adult human population. Nat. Med. 2019;25:242–248. [PubMed] [Google Scholar]

160. Guy J., Feuer W.J., Davis J.L., Porciatti V., Gonzalez P.J., Koilkonda R.D., Yuan H., Hauswirth W.W., Lam B.L. Gene therapy for Leber hereditary optic neuropathy: low- and medium-dose visual results. Ophthalmology. 2017;124:1621–1634. [PMC free article] [PubMed] [Google Scholar]

161. Wan X., Pei H., Zhao M.J., Yang S., Hu W.K., He H., Ma S.Q., Zhang G., Dong X.Y., Chen C. Efficacy and safety of rAAV2-ND4 treatment for Leber’s hereditary optic neuropathy. Sci. Rep. 2016;6:21587. [PMC free article] [PubMed] [Google Scholar]

162. Dalkara D., Byrne L.C., Klimczak R.R., Visel M., Yin L., Merigan W.H., Flannery J.G., Schaffer D.V. Invivo-directed evolution of a new adeno-associated virus for therapeutic outer retinal gene delivery from the vitreous. Sci. Transl. Med. 2013;5:189ra76. [PubMed] [Google Scholar]

163. Feuer W.J., Schiffman J.C., Davis J.L., Porciatti V., Gonzalez P., Koilkonda R.D., Yuan H., Lalwani A., Lam B.L., Guy J. Gene therapy for Leber hereditary optic neuropathy: initial results. Ophthalmology. 2016;123:558–570. [PMC free article] [PubMed] [Google Scholar]

164. Vignal C., Uretsky S., Fitoussi S., Galy A., Blouin L., Girmens J.-F., Bidot S., Thomasson N., Bouquet C., Valero S. Safety of rAAV2/2-ND4 gene therapy for leber hereditary optic neuropathy. Ophthalmology. 2018;125:945–947. [PubMed] [Google Scholar]

165. Bouquet C., Vignal Clermont C., Galy A., Fitoussi S., Blouin L., Munk M.R., Valero S., Meunier S., Katz B., Sahel J.A., Thomasson N. Immune response and intraocular inflammation in patients with Leber hereditary optic neuropathy treated with intravitreal injection of recombinant adeno-associated virus 2 carrying the ND4 gene: a secondary analysis of a phase 1/2 clinical trial. JAMA Ophthalmol. 2019;137:399–406. [PMC free article] [PubMed] [Google Scholar]

166. Heier J.S., Kherani S., Desai S., Dugel P., Kaushal S., Cheng S.H., Delacono C., Purvis A., Richards S., Le-Halpere A. Intravitreous injection of AAV2-sFLT01 in patients with advanced neovascular age-related macular degeneration: a phase 1, open-label trial. Lancet. 2017;390:50–61. [PubMed] [Google Scholar]

167. Mingozzi F., High K.A. Therapeutic invivo gene transfer for genetic disease using AAV: progress and challenges. Nat. Rev. Genet. 2011;12:341–355. [PubMed] [Google Scholar]

168. Collaud F., Bortolussi G., Guianvarc’h L., Aronson S.J., Bordet T., Veron P., Charles S., Vidal P., Sola M.S., Rundwasser S. Preclinical development of an AAV8-hUGT1A1 vector for the treatment of Crigler-Najjar syndrome. Mol. Ther. Methods Clin. Dev. 2018;12:157–174. [PMC free article] [PubMed] [Google Scholar]

169. Vidal P., Pagliarani S., Colella P., Costa Verdera H., Jauze L., Gjorgjieva M., Puzzo F., Marmier S., Collaud F., Simon Sola M. Rescue of GSDIII phenotype with gene transfer requires liver- and muscle-targeted GDE expression. Mol. Ther. 2018;26:890–901. [PMC free article] [PubMed] [Google Scholar]

170. Vilà L., Elias I., Roca C., Ribera A., Ferré T., Casellas A., Lage R., Franckhauser S., Bosch F. AAV8-mediated Sirt1 gene transfer to the liver prevents high carbohydrate diet-induced nonalcoholic fatty liver disease. Mol. Ther. Methods Clin. Dev. 2014;1:14039. [PMC free article] [PubMed] [Google Scholar]

171. D’Avola D., López-Franco E., Sangro B., Pañeda A., Grossios N., Gil-Farina I., Benito A., Twisk J., Paz M., Ruiz J. Phase I open label liver-directed gene therapy clinical trial for acute intermittent porphyria. J.Hepatol. 2016;65:776–783. [PubMed] [Google Scholar]

172. Fuchs, S.P., Martinez-Navio, J.M., Rakasz, E.G., Gao, G., and Desrosiers, R.C. Liver-directed but not muscle-directed AAV-antibody gene transfer limits humoral immune responses in rhesus monkeys. Mol. Ther. Published online November 25, 2019. 10.1016/j.omtm.2019.11.010. [PMC free article] [PubMed]

173. Ruzo A., Garcia M., Ribera A., Villacampa P., Haurigot V., Marcó S., Ayuso E., Anguela X.M., Roca C., Agudo J. Liver production of sulfamidase reverses peripheral and ameliorates CNS pathology in mucopolysaccharidosis IIIA mice. Mol. Ther. 2012;20:254–266. [PMC free article] [PubMed] [Google Scholar]

174. Sorrentino N.C., D’Orsi L., Sambri I., Nusco E., Monaco C., Spampanato C., Polishchuk E., Saccone P., De Leonibus E., Ballabio A., Fraldi A. A highly secreted sulphamidase engineered to cross the blood-brain barrier corrects brain lesions of mice with mucopolysaccharidoses type IIIA. EMBO Mol. Med. 2013;5:675–690. [PMC free article] [PubMed] [Google Scholar]

175. Niemeyer G.P., Herzog R.W., Mount J., Arruda V.R., Tillson D.M., Hathco*ck J., van Ginkel F.W., High K.A., Lothrop C.D., Jr. Long-term correction of inhibitor-prone hemophilia B dogs treated with liver-directed AAV2-mediated factor IX gene therapy. Blood. 2009;113:797–806. [PMC free article] [PubMed] [Google Scholar]

176. Monahan P.E., Sun J., Gui T., Hu G., Hannah W.B., Wichlan D.G., Wu Z., Grieger J.C., Li C., Suwanmanee T. Employing a gain-of-function factor IX variant R338L to advance the efficacy and safety of hemophilia B human gene therapy: preclinical evaluation supporting an ongoing adeno-associated virus clinical trial. Hum. Gene Ther. 2015;26:69–81. [PMC free article] [PubMed] [Google Scholar]

177. Chapin J.R.H., Scheiflinger F., Monahan P.E. An analysis of bleeding rates and factor IX consumption in the phase I/II BAX 335 gene therapy trial in subjects with hemophilia B. Res. Pract. Thromb. Haemost. 2017;1(Suppl 1):144. [Google Scholar]

178. Miesbach W., Meijer K., Coppens M., Kampmann P., Klamroth R., Schutgens R., Tangelder M., Castaman G., Schwäble J., Bonig H. Gene therapy with adeno-associated virus vector 5-human factor IX in adults with hemophilia B. Blood. 2018;131:1022–1031. [PMC free article] [PubMed] [Google Scholar]

179. Von Drygalski A., Giermasz A., Castaman G., Key N.S., Lattimore S., Leebeek F.W.G., Miesbach W., Recht M., Long A., Gut R. Etranacogene dezaparvovec (AMT-061 phase 2b): normal/near normal FIX activity and bleed cessation in hemophilia B. Blood Adv. 2019;3:3241–3247. [PMC free article] [PubMed] [Google Scholar]

180. Spark Therapeutics Spark Therapeutics and Pfizer announce data from 15 participants with hemophilia B showing persistent and sustained factor IX levels with no serious adverse events. 2018. https://www.pfizer.com/news/press-release/press-release-detail/spark_therapeutics_and_pfizer_announce_data_from_15_participants_with_hemophilia_b_showing_persistent_and_sustained_factor_ix_levels_with_no_serious_adverse_events

181. Pipe S., Stine K., Rajasekhar A., Everington T., Poma A., Crombez E. 101HEMB01 is a phase 1/2 open-label, single ascending dose-finding trial of DTX101 (AAVrh10FIX) in patients with moderate/severe hemophilia B that demonstrated meaningful but transient expression of human factor IX (hFIX) Blood. 2017;130(Suppl 1):3331. [Google Scholar]

182. Nathwani A.C., Gray J.T., McIntosh J., Ng C.Y., Zhou J., Spence Y., Cochrane M., Gray E., Tuddenham E.G., Davidoff A.M. Safe and efficient transduction of the liver after peripheral vein infusion of self-complementary AAV vector results in stable therapeutic expression of human FIX in nonhuman primates. Blood. 2007;109:1414–1421. [PMC free article] [PubMed] [Google Scholar]

183. McCarty D.M., Monahan P.E., Samulski R.J. Self-complementary recombinant adeno-associated virus (scAAV) vectors promote efficient transduction independently of DNA synthesis. Gene Ther. 2001;8:1248–1254. [PubMed] [Google Scholar]

184. Bilic I., Monahan P.E., Berg V., Scheiflinger F., Reipert B.M. Whole exome sequencing of patients treated with adeno-associated virus serotype 8-factor IX (AAV8-FIX) gene therapy reveals potential determinants of persistent transgene expression. Res. Pract. Thromb. Haemost. 2019;3(Suppl 1) Abstract OC 31.1, 95. [Google Scholar]

185. Mack D.L., Poulard K., Goddard M.A., Latournerie V., Snyder J.M., Grange R.W., Elverman M.R., Denard J., Veron P., Buscara L. Systemic AAV8-mediated gene therapy drives whole-body correction of myotubular myopathy in dogs. Mol. Ther. 2017;25:839–854. [PMC free article] [PubMed] [Google Scholar]

186. Mendell J.R., Al-Zaidy S., Shell R., Arnold W.D., Rodino-Klapac L.R., Prior T.W., Lowes L., Alfano L., Berry K., Church K. Single-dose gene-replacement therapy for spinal muscular atrophy. N.Engl. J. Med. 2017;377:1713–1722. [PubMed] [Google Scholar]

187. Rangarajan S., Walsh L., Lester W., Perry D., Madan B., Laffan M., Yu H., Vettermann C., Pierce G.F., Wong W.Y., Pasi K.J. AAV5-factor VIII gene transfer in severe hemophilia A. N.Engl. J. Med. 2017;377:2519–2530. [PubMed] [Google Scholar]

188. Herzog R.W., Pierce G.F. Liver gene therapy: reliable and durable? Mol. Ther. 2019;27:1863–1864. [PMC free article] [PubMed] [Google Scholar]

189. Lecomte E., Leger A., Penaud-Budloo M., Ayuso E. Single-stranded DNA virus sequencing (SSV-Seq) for characterization of residual DNA and AAV vector genomes. Methods Mol. Biol. 2019;1950:85–106. [PubMed] [Google Scholar]

190. Margine I., Martinez-Gil L., Chou Y.Y., Krammer F. Residual baculovirus in insect cell-derived influenza virus-like particle preparations enhances immunogenicity. PLoS ONE. 2012;7:e51559. [PMC free article] [PubMed] [Google Scholar]

191. Aguti S., Malerba A., Zhou H. The progress of AAV-mediated gene therapy in neuromuscular disorders. Expert Opin. Biol. Ther. 2018;18:681–693. [PubMed] [Google Scholar]

192. Chamberlain K., Riyad J.M., Weber T. Cardiac gene therapy with adeno-associated virus-based vectors. Curr. Opin. Cardiol. 2017;32:275–282. [PMC free article] [PubMed] [Google Scholar]

193. Ronzitti G., Collaud F., Laforet P., Mingozzi F. Progress and challenges of gene therapy for Pompe disease. Ann. Transl. Med. 2019;7:287. [PMC free article] [PubMed] [Google Scholar]

194. Callejas D., Mann C.J., Ayuso E., Lage R., Grifoll I., Roca C., Andaluz A., Ruiz-de Gopegui R., Montané J., Muñoz S. Treatment of diabetes and long-term survival after insulin and gluco*kinase gene therapy. Diabetes. 2013;62:1718–1729. [PMC free article] [PubMed] [Google Scholar]

195. Mueller C., Chulay J.D., Trapnell B.C., Humphries M., Carey B., Sandhaus R.A., McElvaney N.G., Messina L., Tang Q., Rouhani F.N. Human Treg responses allow sustained recombinant adeno-associated virus-mediated transgene expression. J.Clin. Invest. 2013;123:5310–5318. [PMC free article] [PubMed] [Google Scholar]

196. French R.A., Samelson-Jones B.J., Niemeyer G.P., Lothrop C.D., Jr., Merricks E.P., Nichols T.C., Arruda V.R. Complete correction of hemophilia B phenotype by FIX-Padua skeletal muscle gene therapy in an inhibitor-prone dog model. Blood Adv. 2018;2:505–508. [PMC free article] [PubMed] [Google Scholar]

197. Gernoux G., Wilson J.M., Mueller C. Regulatory and exhausted Tcell responses to AAV capsid. Hum. Gene Ther. 2017;28:338–349. [PMC free article] [PubMed] [Google Scholar]

198. Fraites T.J., Jr., Schleissing M.R., Shanely R.A., Walter G.A., Cloutier D.A., Zolotukhin I., Pauly D.F., Raben N., Plotz P.H., Powers S.K. Correction of the enzymatic and functional deficits in a model of Pompe disease using adeno-associated virus vectors. Mol. Ther. 2002;5:571–578. [PubMed] [Google Scholar]

199. Sun B., Zhang H., Franco L.M., Brown T., Bird A., Schneider A., Koeberl D.D. Correction of glycogen storage disease type II by an adeno-associated virus vector containing a muscle-specific promoter. Mol. Ther. 2005;11:889–898. [PubMed] [Google Scholar]

200. Kay M.A., Manno C.S., Ragni M.V., Larson P.J., Couto L.B., McClelland A., Glader B., Chew A.J., Tai S.J., Herzog R.W. Evidence for gene transfer and expression of factor IX in haemophilia B patients treated with an AAV vector. Nat. Genet. 2000;24:257–261. [PubMed] [Google Scholar]

201. Brantly M.L., Spencer L.T., Humphries M., Conlon T.J., Spencer C.T., Poirier A., Garlington W., Baker D., Song S., Berns K.I. Phase I trial of intramuscular injection of a recombinant adeno-associated virus serotype 2 alphal-antitrypsin (AAT) vector in AAT-deficient adults. Hum. Gene Ther. 2006;17:1177–1186. [PubMed] [Google Scholar]

202. Brantly M.L., Chulay J.D., Wang L., Mueller C., Humphries M., Spencer L.T., Rouhani F., Conlon T.J., Calcedo R., Betts M.R. Sustained transgene expression despite T lymphocyte responses in a clinical trial of rAAV1-AAT gene therapy. Proc. Natl. Acad. Sci. USA. 2009;106:16363–16368. [PMC free article] [PubMed] [Google Scholar]

203. Flotte T.R., Trapnell B.C., Humphries M., Carey B., Calcedo R., Rouhani F., Campbell-Thompson M., Yachnis A.T., Sandhaus R.A., McElvaney N.G. Phase 2 clinical trial of a recombinant adeno-associated viral vector expressing α1-antitrypsin: interim results. Hum. Gene Ther. 2011;22:1239–1247. [PMC free article] [PubMed] [Google Scholar]

204. Stroes E.S., Nierman M.C., Meulenberg J.J., Franssen R., Twisk J., Henny C.P., Maas M.M., Zwinderman A.H., Ross C., Aronica E. Intramuscular administration of AAV1-lipoprotein lipase S447X lowers triglycerides in lipoprotein lipase-deficient patients. Arterioscler. Thromb. Vasc. Biol. 2008;28:2303–2304. [PubMed] [Google Scholar]

205. Gaudet D., Méthot J., Déry S., Brisson D., Essiembre C., Tremblay G., Tremblay K., de Wal J., Twisk J., van den Bulk N. Efficacy and long-term safety of alipogene tiparvovec (AAV1-LPLS447X) gene therapy for lipoprotein lipase deficiency: an open-label trial. Gene Ther. 2013;20:361–369. [PMC free article] [PubMed] [Google Scholar]

206. Ferreira V., Twisk J., Kwikkers K., Aronica E., Brisson D., Methot J., Petry H., Gaudet D. Immune responses to intramuscular administration of alipogene tiparvovec (AAV1-LPLS447X) in a phase II clinical trial of lipoprotein lipase deficiency gene therapy. Hum. Gene Ther. 2014;25:180–188. [PMC free article] [PubMed] [Google Scholar]

207. Carpentier A.C., Frisch F., Labbé S.M., Gagnon R., de Wal J., Greentree S., Petry H., Twisk J., Brisson D., Gaudet D. Effect of alipogene tiparvovec (AAV1-LPLS447X) on postprandial chylomicron metabolism in lipoprotein lipase-deficient patients. J.Clin. Endocrinol. Metab. 2012;97:1635–1644. [PubMed] [Google Scholar]

208. Ferreira V., Petry H., Salmon F. Immune responses to AAV-vectors, the Glybera example from bench to bedside. Front. Immunol. 2014;5:82. [PMC free article] [PubMed] [Google Scholar]

209. Mays L.E., Wang L., Lin J., Bell P., Crawford A., Wherry E.J., Wilson J.M. AAV8 induces tolerance in murine muscle as a result of poor APC transduction, Tcell exhaustion, and minimal MHCI upregulation on target cells. Mol. Ther. 2014;22:28–41. [PMC free article] [PubMed] [Google Scholar]

210. Burzyn D., Kuswanto W., Kolodin D., Shadrach J.L., Cerletti M., Jang Y., Sefik E., Tan T.G., Wagers A.J., Benoist C., Mathis D. A special population of regulatory Tcells potentiates muscle repair. Cell. 2013;155:1282–1295. [PMC free article] [PubMed] [Google Scholar]

211. Le Guiner C., Servais L., Montus M., Larcher T., Fraysse B., Moullec S., Allais M., François V., Dutilleul M., Malerba A. Long-term microdystrophin gene therapy is effective in a canine model of duch*enne muscular dystrophy. Nat. Commun. 2017;8:16105. [PMC free article] [PubMed] [Google Scholar]

212. Lowes L.P., Alfano L.N., Arnold W.D., Shell R., Prior T.W., McColly M., Lehman K.J., Church K., Sproule D.M., Nagendran S. Impact of age and motor function in a phase 1/2A study of infants with SMA type 1 receiving single-dose gene replacement therapy. Pediatr. Neurol. 2019;98:39–45. [PubMed] [Google Scholar]

213. Al-Zaidy S., Pickard A.S., Kotha K., Alfano L.N., Lowes L., Paul G., Church K., Lehman K., Sproule D.M., Dabbous O. Health outcomes in spinal muscular atrophy type 1 following AVXS-101 gene replacement therapy. Pediatr. Pulmonol. 2019;54:179–185. [PMC free article] [PubMed] [Google Scholar]

214. Hoy S.M. Onasemnogene Abeparvovec: First Global Approval. Drugs. 2019;79:1255–1262. [PubMed] [Google Scholar]

215. Hinderer C., Katz N., Buza E.L., Dyer C., Goode T., Bell P., Richman L.K., Wilson J.M. Severe Toxicity in Nonhuman Primates and Piglets Following High-Dose Intravenous Administration of an Adeno-Associated Virus Vector Expressing Human SMN. Hum. Gene Ther. 2018;29:285–298. [PMC free article] [PubMed] [Google Scholar]

216. Cure SMA AveXis issues community statement on AVXS-101 clinical hold. 2019. https://www.curesma.org/avexis-community-statement-clinical-hold2019/

217. Piguet F., Alves S., Cartier N. Clinical gene therapy for neurodegenerative diseases: past, present, and future. Hum. Gene Ther. 2017;28:988–1003. [PubMed] [Google Scholar]

218. Gray S.J., Nagabhushan Kalburgi S., McCown T.J., Jude Samulski R. Global CNS gene delivery and evasion of anti-AAV-neutralizing antibodies by intrathecal AAV administration in non-human primates. Gene Ther. 2013;20:450–459. [PMC free article] [PubMed] [Google Scholar]

219. Haurigot V., Marcó S., Ribera A., Garcia M., Ruzo A., Villacampa P., Ayuso E., Añor S., Andaluz A., Pineda M. Whole body correction of mucopolysaccharidosis IIIA by intracerebrospinal fluid gene therapy. J.Clin. Invest. 2013;123:3254–3271. [PMC free article] [PubMed] [Google Scholar]

220. Brown G.C., Neher J.J. Microglial phagocytosis of live neurons. Nat. Rev. Neurosci. 2014;15:209–216. [PubMed] [Google Scholar]

221. Tufail Y., Cook D., Fourgeaud L., Powers C.J., Merten K., Clark C.L., Hoffman E., Ngo A., Sekiguchi K.J., O’Shea C.C. Phosphatidylserine exposure controls viral innate immune responses by microglia. Neuron. 2017;93:574–586.e8. [PMC free article] [PubMed] [Google Scholar]

222. Kierdorf K., Prinz M. Factors regulating microglia activation. Front. Cell. Neurosci. 2013;7:44. [PMC free article] [PubMed] [Google Scholar]

223. Vandamme C., Adjali O., Mingozzi F. Unraveling the complex story of immune responses to AAV vectors trial after trial. Hum. Gene Ther. 2017;28:1061–1074. [PMC free article] [PubMed] [Google Scholar]

224. Samaranch L., Pérez-Cañamás A., Soto-Huelin B., Sudhakar V., Jurado-Arjona J., Hadaczek P., Ávila J., Bringas J.R., Casas J., Chen H. Adeno-associated viral vector serotype 9-based gene therapy for Niemann-Pick disease type A. Sci. Transl. Med. 2019;11:eaat3738. [PMC free article] [PubMed] [Google Scholar]

225. McPhee S.W., Janson C.G., Li C., Samulski R.J., Camp A.S., Francis J., Shera D., Lioutermann L., Feely M., Freese A., Leone P. Immune responses to AAV in a phase I study for Canavan disease. J.Gene Med. 2006;8:577–588. [PubMed] [Google Scholar]

226. Mingozzi F., Anguela X.M., Pavani G., Chen Y., Davidson R.J., Hui D.J., Yazicioglu M., Elkouby L., Hinderer C.J., Faella A. Overcoming preexisting humoral immunity to AAV using capsid decoys. Sci. Transl. Med. 2013;5:194ra92. [PMC free article] [PubMed] [Google Scholar]

227. Ogden P.J., Kelsic E.D., Sinai S., Church G.M. Comprehensive AAV capsid fitness landscape reveals a viral gene and enables machine-guided design. Science. 2019;366:1139–1143. [PMC free article] [PubMed] [Google Scholar]

228. Tse L.V., Klinc K.A., Madigan V.J., Castellanos Rivera R.M., Wells L.F., Havlik L.P., Smith J.K., Agbandje-McKenna M., Asokan A. Structure-guided evolution of antigenically distinct adeno-associated virus variants for immune evasion. Proc. Natl. Acad. Sci. USA. 2017;114:E4812–E4821. [PMC free article] [PubMed] [Google Scholar]

229. Gao G., Wang Q., Calcedo R., Mays L., Bell P., Wang L., Vandenberghe L.H., Grant R., Sanmiguel J., Furth E.E., Wilson J.M. Adeno-associated virus-mediated gene transfer to nonhuman primate liver can elicit destructive transgene-specific Tcell responses. Hum. Gene Ther. 2009;20:930–942. [PMC free article] [PubMed] [Google Scholar]

230. Nguyen D.H., Hurtado-Ziola N., Gagneux P., Varki A. Loss of Siglec expression on T lymphocytes during human evolution. Proc. Natl. Acad. Sci. USA. 2006;103:7765–7770. [PMC free article] [PubMed] [Google Scholar]

231. Li C., Hirsch M., Asokan A., Zeithaml B., Ma H., Kafri T., Samulski R.J. Adeno-associated virus type 2 (AAV2) capsid-specific cytotoxic T lymphocytes eliminate only vector-transduced cells coexpressing the AAV2 capsid invivo. J.Virol. 2007;81:7540–7547. [PMC free article] [PubMed] [Google Scholar]

232. Li H., Murphy S.L., Giles-Davis W., Edmonson S., Xiang Z., Li Y., Lasaro M.O., High K.A., Ertl H.C. Pre-existing AAV capsid-specific CD8+ Tcells are unable to eliminate AAV-transduced hepatocytes. Mol. Ther. 2007;15:792–800. [PubMed] [Google Scholar]

233. Wang L., Figueredo J., Calcedo R., Lin J., Wilson J.M. Cross-presentation of adeno-associated virus serotype 2 capsids activates cytotoxic Tcells but does not render hepatocytes effective cytolytic targets. Hum. Gene Ther. 2007;18:185–194. [PubMed] [Google Scholar]

234. Martino A.T., Basner-Tschakarjan E., Markusic D.M., Finn J.D., Hinderer C., Zhou S., Ostrov D.A., Srivastava A., Ertl H.C., Terhorst C. Engineered AAV vector minimizes invivo targeting of transduced hepatocytes by capsid-specific CD8+ Tcells. Blood. 2013;121:2224–2233. [PMC free article] [PubMed] [Google Scholar]

235. Ertl H.C.J. Preclinical models to assess the immunogenicity of AAV vectors. Cell. Immunol. 2019;342:103722. [PubMed] [Google Scholar]

236. Shoenfeld Y., Gallant L.A., Shaklai M., Livni E., Djaldetti M., Pinkhas J. Gaucher’s disease: a disease with chronic stimulation of the immune system. Arch. Pathol. Lab. Med. 1982;106:388–391. [PubMed] [Google Scholar]

Articles from Molecular Therapy are provided here courtesy of The American Society of Gene & Cell Therapy

AAV Vector Immunogenicity in Humans: A Long Journey to Successful Gene Transfer (2024)
Top Articles
Latest Posts
Article information

Author: Greg Kuvalis

Last Updated:

Views: 6303

Rating: 4.4 / 5 (75 voted)

Reviews: 82% of readers found this page helpful

Author information

Name: Greg Kuvalis

Birthday: 1996-12-20

Address: 53157 Trantow Inlet, Townemouth, FL 92564-0267

Phone: +68218650356656

Job: IT Representative

Hobby: Knitting, Amateur radio, Skiing, Running, Mountain biking, Slacklining, Electronics

Introduction: My name is Greg Kuvalis, I am a witty, spotless, beautiful, charming, delightful, thankful, beautiful person who loves writing and wants to share my knowledge and understanding with you.