Next Article in Journal
The Genomic Characteristics of an Arthritis-Causing Salmonella pullorum
Next Article in Special Issue
A Comprehensive Study of the Effects by Sequence Truncation within Inverted Terminal Repeats (ITRs) on the Productivity, Genome Packaging, and Potency of AAV Vectors
Previous Article in Journal
Characterisation of Soil Bacterial Communities That Exhibit Chemotaxis to Root Exudates from Phosphorus-Limited Plants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

AAV Immunotoxicity: Implications in Anti-HBV Gene Therapy

1
Wits/SAMRC Antiviral Gene Therapy Research Unit, Infectious Diseases and Oncology Research Institute (IDORI), Faculty of Health Sciences, University of the Witwatersrand, Parktown 2193, South Africa
2
Medical Biotechnology and Immunotherapy Research Unit, Institute of Infectious Disease and Molecular Medicine, Faculty of Health Sciences, University of Cape Town, Cape Town 7700, South Africa
3
South African Research Chair in Cancer Biotechnology, Department of Integrative Biomedical Sciences, Faculty of Health Sciences, University of Cape Town, Cape Town 7700, South Africa
*
Author to whom correspondence should be addressed.
Microorganisms 2023, 11(12), 2985; https://doi.org/10.3390/microorganisms11122985
Submission received: 3 November 2023 / Revised: 30 November 2023 / Accepted: 11 December 2023 / Published: 14 December 2023
(This article belongs to the Special Issue Adeno-Associated Virus Biology and AAV Vector-Mediated Gene Therapy)

Abstract

:
Hepatitis B virus (HBV) has afflicted humankind for decades and there is still no treatment that can clear the infection. The development of recombinant adeno-associated virus (rAAV)-based gene therapy for HBV infection has become important in recent years and research has made exciting leaps. Initial studies, mainly using mouse models, showed that rAAVs are non-toxic and induce minimal immune responses. However, several later studies demonstrated rAAV toxicity, which is inextricably associated with immunogenicity. This is a major setback for the progression of rAAV-based therapies toward clinical application. Research aimed at understanding the mechanisms behind rAAV immunity and toxicity has contributed significantly to the inception of approaches to overcoming these challenges. The target tissue, the features of the vector, and the vector dose are some of the determinants of AAV toxicity, with the latter being associated with the most severe adverse events. This review discusses our current understanding of rAAV immunogenicity, toxicity, and approaches to overcoming these hurdles. How this information and current knowledge about HBV biology and immunity can be harnessed in the efforts to design safe and effective anti-HBV rAAVs is discussed.

1. Introduction

Viral infections caused by “old” viruses, such as hepatitis B virus (HBV), and epidemics or pandemics caused by “new” viruses, such as severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), also known as COVID-19, continue to threaten human existence. It is estimated that 296 million individuals are chronically infected with HBV, the etiological agent of hepatitis B. Chronic hepatitis B (CHB) results in about one million deaths annually owing to HBV-related liver complications such as cirrhosis and hepatocellular carcinoma (HCC) [1].
Vaccination remains the primary mechanism to reduce HBV-related mortality. The first available vaccines targeting HBV were plasma-derived and were produced by harvesting the subvirion particles of the HBV surface antigen (HBsAg) from the plasma of chronic HBV asymptomatic donors. In 1986, second-generation HBV vaccines, consisting of yeast-derived recombinant HBsAg, replaced plasma-derived vaccines. Currently, recombinant DNA vaccines, composed of highly purified HBsAg adsorbed to an aluminum adjuvant, are being used for universal HBV immunization [2,3]. After three doses of the HBV vaccine, a long-term immune response can be achieved in most immune-competent individuals. However, 5–10% of those who are vaccinated are non-responders. A protective response is defined as anti-HBsAg antibody levels of ≥100 IU/L; however, in the case of vaccine ‘non-responders’, their anti-HBsAg antibody levels are lower than 10 IU/L [4,5,6]. While the molecular mechanisms responsible for non-responsiveness remain largely unclear, some may include the failure of antigen presentation and insufficient production of T helper subset 1 (Th1) and 2 (Th2) cytokines upon vaccination. Several known factors contributing to non-responsiveness include certain genetic risk factors, a compromised immune response, chronic diseases such as diabetes, and being older than 30 years [7,8]. Efforts to develop novel vaccines include the use of RNA- and viral vector-based approaches. RNA lipid nanoparticle (LNP) formulations and adenovirus (AdV)-based vaccines have performed well in pre-clinical studies [9,10,11]. Recombinant adeno-associated viral vectors (rAAVs) have also been explored for vaccine development. However, as a result of the inherent low immune stimulation, rAAVs are not favored for vaccine development [12,13]. Nevertheless, they are promising candidates for antiviral immunotherapy [14,15].
The fact that anti-HBV drugs, comprising nucleoside/nucleotide analogs and immune modulators, do not provide a sterilizing cure emphasizes the need for innovative ways of designing anti-HBV therapies. Extensive research into using gene therapy for the treatment of HBV infection reported promising outcomes in pre-clinical studies. These include RNA interference (RNAi)-based therapies and engineered nucleases such as zinc finger nucleases (ZFNs), transcription activator-like effector nucleases (TALENs), and the RNA-guided clustered regulatory interspaced short palindromic repeats (CRISPR) and CRISPR-associated (Cas) systems [16,17,18,19]. However, finding safe and effective gene delivery vehicles for these nucleic acid sequences remains one of the toughest hurdles in anti-HBV gene therapy. To target CHB, long-term expression of anti-HBV effectors is essential. Whereas non-viral vectors such as lipid-based nano-carrier formulations have several attractive features, they result in transient therapeutic effects [20,21]. The features of rAAVs that have made them one of the leading gene delivery platforms in recent years include rare integration into the host genome, high transduction efficiencies in multiple tissues, including the liver, and sustained transgene expression. The existence of multiple serotypes and artificially synthesized variants gives AAV-derived vectors versatility and several candidates have now been approved for clinical use.
Glybera was the first AAV-based drug to be approved for clinical application in 2012, which was for the treatment of lipoprotein lipase deficiency. The recent approval of Luxturna, Zolgensma, Hemgenix, Elevidys, and Roctavian for the treatment of retinal dystrophy, spinal muscular atrophy, hemophilia B, Duchenne’s muscular dystrophy and hemophilia A, respectively, further highlight the potential of AAV-based therapeutics. Currently, there are over 130 clinical trials using rAAVs and the feasibility of using AAVs for HBV infection treatment has been demonstrated in pre-clinical studies [16,22,23]. However, high AAV immunogenicity and the resultant toxicity are becoming apparent in non-human primates, as shown in clinical studies [24].
Although AAVs elicit a modest innate immune response when compared to other viral vectors, this is enough to aggravate the adaptive immune response and cause toxicity. A high vector dose is a key factor in most of the studies, which showed AAV-induced toxicities, as demonstrated when using doses higher than 1 × 1014 vg/kg to treat neuromuscular disease. The death of four patients who received 1.3 × 1014 vg/kg or 3.5 × 1014 vg/kg for X-linked myotubular myopathy in a clinical trial (NCT03199469) represents the highest number of deaths reported in a single AAV clinical trial [25]. Although more details are required, the recent deaths of two patients increased the number of deaths in patients who received Zolgensma to three, and further raised safety concerns [26]. In a recent FDA-approved clinical trial for the treatment of Duchenne’s muscular dystrophy (DMD), the administration of a high dose (1 × 1014 vg/kg) of rAAV containing a dead Staphylococcus aureus Cas 9 (dSaCas9) led to the death of one patient [27]. Several studies have implicated vector-specific immune responses as the major cause of most of these fatalities [24,28,29]. Extensive research has recently focused on understanding AAV immunotoxicity and finding ways to overcome this challenge. Methods adopted include manipulating the host’s immune system or modifying the AAVs to evade immune detection. This review summarizes current knowledge on rAAV immunogenicity and toxicity, discusses progress in overcoming these challenges, and explores how these developments and current knowledge on HBV biology and immunity can benefit endeavors in rAAV-based anti-HBV therapy development.

2. Biology of HBV

HBV is genomically diverse, with 10 genotypes (A-J) and various sub-genotypes [30]. The virion genome is located within an icosahedral capsid, which, in turn, is encased in a lipid membrane to constitute a spherical viral particle of 42 nm. The genome is made up of 3.2 kb of relaxed circular DNA (rcDNA) with partially formed positive and complete negative strands. On the 5′ positive strand is a pre-genomic messenger RNA (pg-mRNA) fragment serving a primer function during positive DNA strand repair. The genome carries four overlapping open reading frames (ORFs): pre-core/Core (pre-c/C), Pol, pre-surface/Surface (pre-s/S), and X. These ORFs bear seven translation initiation codons. Pre-core/C encodes the hepatitis B core (HBcAg), which forms the viral capsid, along with a secreted hepatitis B e-antigen (HBeAg). Pol encodes viral polymerase, while pre-s/S makes three envelope structural proteins, referred to by size: large (L-HBs), middle (M-HBs), and small (S-HBs) surface antigens. X produces the HBx protein, which is critical for the episomal maintenance of the viral genome. HBV entry and genome transfer into the nucleus of hepatocytes is initiated by binding to the heparan sulfate proteoglycans (HSPGs) at the cell surface. Virion uptake by binding to a receptor, the sodium taurocholate co-transporting peptide (NTCP), in an endocytosis-dependent manner follows. Infected hepatocytes repair the partly double-stranded genome in the nucleus and produce stable covalently closed circular DNA (cccDNA), which serves as a template for viral gene expression. From the cccDNA, pre-genomic RNA (3.5 Kb), viral surface RNAs (2.4 Kb and 2.1 Kb), and HBx RNA (0.7 Kb) are produced. These are required for viral replication, capsid formation, episomal maintenance, and infectivity [31].

3. Immune Response to HBV

The host immune responses not only control the spread of HBV but also contribute to liver inflammation and damage while fighting off the infection. In the early stages of HBV infection, innate immune responses are triggered by viral nucleic acids and viral proteins, leading to the production of antiviral cytokines (e.g., IFN-α) and natural killer (NK) cell activation to impede the initial spread of HBV [32]. The innate immune response is essential for the activation of adaptive immunity. While innate immunity is important for the rapid control of viral replication, the presence of an increase in the number and strength of HBV-specific T (e.g., cytotoxic or helper T cells) and B cells during infection is the ultimate determinant of the outcome of infection. The production of cytokines by CD4+ T cells is necessary for the development of CD8+ cytotoxic T lymphocytes, which clear HBV infection by destroying the infected hepatocytes [33,34].
CHB is often associated with poor effector T-cell responses. This immune ‘exhaustion’ is characterized by poor cytotoxic T-cell activity and impaired cytokine production [35]. The role of regulatory T cells (Tregs) in this ‘immune exhaustion’ has been widely demonstrated [36,37,38,39]. In a poorly understood mechanism, HBV infection induces Treg responses [40]. Recent studies have suggested the involvement of host factors such as heat shock protein 60, furin protein, and viral proteins such as HBeAg and HBx [40,41,42,43]. The study of anti-HBV adaptive immunity has primarily focused on the T-cell responses; however, the various immune functions of B cells, including antigen presentation, antibody secretion, and immune regulation mediate HBV persistence, tolerance, and liver damage [44].

4. Gene Therapies against HBV

The inhibition of HBV replication by the exogenously supplied mimics of RNAi intermediates, such as artificial primary microRNA (apri-miRNA), short hairpin RNA (shRNA), and short interfering RNAs (siRNA), has been clearly demonstrated. These RNAi activators recruit target HBV RNA to the RNA-induced silencing complex (RISC) for degradation (see Figure 1) [17,20,45,46]. The highly stable replication-competent cccDNA minichromosome persists in infected cells for an extended time and is the major contributor to relapse following treatment withdrawal. Hence, using designer nucleases to inactivate cccDNA permanently has become an attractive option for novel anti-HBV gene therapy development. The engineered nucleases are similar, in the sense that they comprise a DNA binding domain and a nuclease domain. The DNA binding domain in the case of HBV binds to cccDNA or rcDNA, whereas the nuclease domain creates site-specific DNA double-stranded breaks. This recruits repair machinery and, more often, non-homologous end-joining is favored over homology-directed repair, due to the lack of a donor sequence (see Figure 1) [47].
ZFNs typically consist of three to six ZFs that are capable of binding to a DNA sequence of around 9–18 bp [48]. Cleavage of the target DNA is dependent on two ZFN monomers to form an active nuclease (see Figure 1). ZFNs are not commonly used in the field of anti-HBV gene therapy due to the difficulty of constructing them. Two studies have reported on the use of ZFNs against HBV [49,50]. While reduction of viral replication markers was noted, targeted mutagenesis against cccDNA was not present. TALENs differ slightly from ZFNs as they comprise 17.5 tandem repeats comprising 34 amino acids. The amino acids at positions 12 and 13 are referred to as repeat variable diresidues, which are capable of recognizing the signal nucleotides of target DNA. Similar to ZFNs, TALENs function as pairs for site-directed double-stranded DNA cleavage to occur (see Figure 1). Studies using anti-HBV TALENs demonstrated the targeted disruption of HBV DNA [51,52]. To improve specificity, second-generation TALENs make use of mutated FokI nuclease domains as obligate heterodimeric pairs. A study utilizing second-generation anti-HBV TALENs reported the same effectiveness at targeting cccDNA as first-generation TALENs, with improved specificity [53].
CRISPR consists of two components involving single-guide RNA (sgRNA), which forms a complex with the Cas 9 protein to direct cleavage at the target site [54]. The sgRNA is composed of CRISPR-RNA (crRNA) and trans-activating crRNA (tracrRNA) and guides the Cas 9 nuclease to the target site (see Figure 1). The most common Cas 9 nucleases used are taken from Streptococcus pyogenes (S. pyogenes) and Staphylococcus aureus (S. aureus) [55,56]. CRISPR/Cas 9 has been at the forefront of HBV gene therapy, with numerous studies showcasing its ability to target cccDNA. Several studies have demonstrated significant cccDNA cleavage and reduction of circulating HBV particles in vitro and/or in mouse models [18,57,58,59,60,61,62,63,64]. While off-target cleavage and immunity against the Cas protein is a concern, there is significant progress in developing CRISPR/Cas systems with improved specificity, designing methods to detect off-target effects, and identifying Cas or Cas-like proteins with low seroprevalence [65,66,67,68,69,70]. The progression of EBT-101, a CRISPR/Cas therapeutic against HIV in the Phase I/II clinical trial, highlights the potential of CRISPR/Cas therapeutics against viral infections (https://clinicaltrials.gov/study/NCT05144386, accessed on the 24 November 2023).

5. Biology of AAVs

AAVs are small, non-enveloped viruses belonging to the Dependoparvovirus genus of the Parvoviridae family of viruses, in which the AAV serotype 2 (AAV2) serves as a prototype. Encapsulated in the ~26 nm capsid is a single-stranded genome of about 4.7 kb, with inverted terminal repeats (ITR) flanking rep and cap ORFs. Embedded within the cap ORF is the assembly activation protein-encoding sequence. The ITRs contain the cis-acting Rep protein binding element (RBE), a terminal resolution site (TRS), and three pairs of complementary palindromes (A-A’, B-B’, and C-C’), as well as a single non-palindromic region (D). The palindromes allow for the genome’s T-shaped hairpin secondary structure to form, either in a flip or flop configuration. The ITRs’ cis-acting elements and TRS are essential for genome replication, packaging, chromosomal integration, rescue, and circularization, among other functions. A trio of promoters (p40, p5, and p19) drives gene expression from the rep and cap ORFs. The Rep78, Rep68, Rep52, and Rep40 proteins encoded by rep ORF are named according to their molecular size and serve different functions during genome replication and packaging. Cap ORF encodes the capsid proteins, VP1, VP2, and VP3 (see Figure 2) [71,72,73].
The high tropic heterogeneity of AAV capsids, with 13 known AAV serotypes and several variants, is mediated by their different primary receptors. Moreover, there is a handful of secondary receptors that co-mediate transduction and infectivity by AAVs [74,75]. The AAVs’ replicative incompetence leads to viral latency post-infection and genome integration at a preferential site, AAVS1, within the host genome [76]. However, AAVs benefit from helper gene expression by adenovirus (Ad) or the herpes simplex virus (HSV), which enables productive infection [77].

6. AAV-Derived Vectors and Their Application in Anti-HBV Gene Therapy

Besides tropism, the transgene cassette size, duration, and rate of expression need to be considered when designing a viral vector for therapeutic application. The ability of AAVs to form stable, episomal, and circular or concatemeric genomes in the nucleus enables AAV persistence and long-lasting gene expression [78]. All rAAVs have been developed as vectors for gene therapy by deleting both the rep and cap ORFs, thus maintaining only the ITRs and allowing the insertion of transgenes (see Figure 2). The deleted sequences and helper sequences are supplied in trans during vector production. The single-stranded AAV (ssAAV) and self-complementary AAV (scAAV) constructs used in various studies mimic the genome conformations that occur during the AAV replication cycle. Although the construction of ssAAVs only involves replacing the viral genes, scAAVs also require the mutation of the TRS to abolish nearby ITR resolution [79]. scAAVs thus carry at least three T-shaped hairpin loops (see Figure 2). Compared to scAAVs, ssAAVs have twice the transgene capacity. ssAAVs can accommodate cassettes of about 4.7 kb, but one drawback is their delayed transgene expression compared to scAAVs. This is the result of a requirement for a rate-limiting second-strand synthesis step prior to transgene expression [80]. scAAVs are designed to have a ready-to-express double-stranded conformation, which results in more efficient transduction and faster transgene expression kinetics [81].
The scAAV-mediated delivery of anti-HBV shRNA or apri-miRNA has been extensively explored in preclinical studies, with promising outcomes. Targeting the S or X ORFs using scAAVs expressing shRNA or apri-miRNA in preclinical studies resulted in the inhibition of HBV replication by up to ~98% [22,46,82,83]. As a result of difficulties with designing ZFNs and the large size of TALENs, research on their delivery using AAVs has been limited. One study demonstrated rcDNA-targeted mutagenesis and a reduction in HBV viral particles when using scAAVs expressing anti-HBV ZFNs against HBV P, C, or X genes in cultured cells. However, cytotoxicity was observed with one of the ZFNs [50]. Although the delivery of TALENs targeting HBV has not been reported, a dual-AAV approach with ssAAVs is an option (see Figure 2).
An all-in-one ssAAV comprising a saCas9 and sgRNA targeting S or P ORF resulted in the efficient inactivation of HBV replication and the mutation of cccDNA in HepG2-hNTCP cells, a liver-derived HepG2 cell line overexpressing hNTCP that is infectable by HBV [16]. A similar study utilized sgRNAs targeting all HBV ORFs and demonstrated a decrease in HBsAg and HBeAg in vitro [84]. Many studies have now focused on the feasibility of using CRISPR/Cas 9 in vivo and have shown promising results. CRISPR/Cas 9 targeting the C and X ORF resulted in a significant reduction of HBV DNA in mice [19,85,86] (see Figure 2). Humanized mice can produce cccDNA and present a useful model of HBV infection. The demonstration that the AAV-mediated expression of HBV-targeting sgRNAs can reduce cccDNA in humanized mice further reinforced the suitability of AAVs for anti-HBV gene editor delivery [18].

7. Toxicity of rAAV

Features within both the AAV genome and capsids contribute to AAV toxicity. Certain cis-regulatory elements found in commonly used non-tissue specific promoters, such as cytomegalovirus (CMV) and chicken beta actin (CAG) promoters, p53 binding elements within the ITRs, and ITR promoter activity have been linked to toxicity [87,88,89,90]. rAAV integration-mediated oncogenicity has been controversial, with contradicting information. However, it is clear that this is a rare event and seems to be dominant in neonatal mice injected with high doses of rAAVs carrying strong promoters [91,92,93]. The majority of these integrations occurred within the microRNA (miR)-341-encoding region within the Rian locus of mice, while humans have no ortholog of this miR. Evidence of AAV integration in humans has been of the wild-type AAV integration in the liver. These integrations were also enriched in non-HCC tissues, suggesting that they are not major contributors to HCC in humans [94]. Although genotoxicity has not been demonstrated in long-term AAV-based gene therapy clinical studies, a demonstration that certain enhancer–promoter elements can also facilitate rAAV integration in the human liver will necessitate the careful selection of promoters to drive transgene expression [76]. Whereas genotoxicity and cytotoxicity are a challenge, immunotoxicity appears to be the major obstacle to successful clinical studies and will be the focus of this review.

8. Immune Response to rAAV

The initial, short-lived, and non-antigen-specific innate immune response to AAVs is most commonly initiated by the interaction between the pathogen-associated molecular patterns (PAMPs) of the AAV capsid or the genome (e.g., unmethylated CpG islands in the dsDNA) and the host’s pathogen recognition receptors (PRR), such as the Toll-like receptors (TLRs) on the cell surface (e.g., TLR9, which binds to CpGs, and TLR2, which binds to the AAV antigen). PAMP–PRR interaction results in the activation of transcription factors such as nuclear factor kappa B (NF-κB) and interferon regulatory factors. This mediates the expression of inflammatory cytokines and chemokines, the maturation of dendritic cells into antigen-presenting cells (APCs), and the activation of innate immune cells such as macrophages, NK cells, and neutrophils [28,95].
In concert with adaptive immune response factors, complement system activation has been observed in patients who received AAVs [96]. The complement pathway is a component of the innate immune system, which comprises about 50 plasma proteins that interact to activate the inflammatory responses and target pathogens for killing [97]. Although not well understood, AAV binding to antibodies results in the activation of the complement system. This results in the direct killing or neutralization of the AAVs, inflammation, and tissue injury [98,99]. The induction of endoplasmic reticulum stress by AAV transgene products, such as Factor VIII, which is expressed from strong promoters, can also activate an inflammatory response via a poorly understood mechanism [100,101].
The prolonged humoral adaptive immune response to AAVs is initiated by the recognition of the AAV capsid by B cell receptors on naïve or memory B cells, which is followed by antibody secretion. Naïve B cells can also differentiate into short-lived plasma cells that secrete immature antibodies. Alternatively, T helper cells can mediate B cell maturation and differentiation into long-lived plasma cells and memory B cells. Upon antigen re-encounter, these cells secrete high-affinity antibodies, which can then bind to the AAV capsid, resulting in the opsonization and removal of immune complexes by macrophages. B cell responses against capsids do not typically induce tissue damage, but they can neutralize the vector upon re-administration [102].
The cellular immune response to AAVs is initiated by T cell activation by peptides from the intracellular degradation of AAV proteins. The peptides bind to MHC-associated molecules before their translocation to the cell surface for presentation to CD8+ T cells (MHC I) or CD4+ T cells (MHC II). Activated CD4+ T cells can differentiate into several Th cell subsets, e.g., Th1 cells differentiate to promote CD8+ T cell responses, Th2 cells differentiate to drive B cell activation, and Th17 cells differentiate to activate neutrophils or regulatory T cells to dampen T cell responses. Activated CD8+ T cells release cytokines such as IFN-ɣ or lyse APCs by releasing perforin and granzyme B [103,104]. To maintain the balance between self-tolerance and autoimmunity, Tregs suppress the proliferation and/or effector functions of CD4+ T cells, CD8+ T cells, B cells, NK cells, and antigen-presenting cells (APCs) [105,106,107].

9. Immune Response to AAVs in the Liver

The liver contains the largest population of NK cells and macrophages, called Kupfer cells (KCs), which directly clear foreign molecules that enter the portal circulation through the gut. However, unlike other macrophages, in most cases, KCs remain in an anti-inflammatory state and are poor activators of T cells. NK cells can serve both immunosuppressive and immune-activating functions through their inhibitory and activating receptors [108,109]. It is well established that liver Tregs mediate immune tolerance via several antigen-specific and non-specific mechanisms, which include the suppression of effector T cells, B cells, and APCs [109]. Programmed death ligand-1 (PD-L1)-expressing and IL-10-producing KCs drive Treg expansion and mediate the conversion of antigen-specific effector T cells into Tregs (see Figure 3) [110,111].
Hepatocytes comprise 80% of the cells in the liver and carry out diverse functions that include protein synthesis, metabolism, and detoxification. Hepatocytes have antigen-presenting capabilities, and they can interact directly with T cells. It is clear that under steady-state conditions, hepatocytes do not express MHC-II. Interestingly, in chronic liver diseases such as hepatitis, hepatocytes express higher levels of MHC-II molecules and induce the expression of Foxp3 in CD4+ T cells (see Figure 3) [110,112,113,114]. In non-parenchymal hepatocytes, the sensing of AAV particles by constitutive TLR2 resulted in aggravated innate responses [115]. Other liver-regulatory cells implicated in modulating immune tolerance include type 1 regulatory T cells (Tr1), regulatory B cells, IL-10-producing liver sinusoidal endothelial cells (LSECs), and IL-10-producing dendritic cells [109].
The fact that AAVs can exploit the tolerogenic nature of the liver makes them attractive for liver-directed gene transfer [109]. AAV liver gene transfer can result in the deletion of antigen-specific effector T cells and the induction of Treg activation and expansion, leading to the suppression of transgene product specific effector T cell responses and antibody production [116,117]. Although the mechanisms are not well understood, a demonstration that AAV administration to the liver results in IL-10 production by KCs in response to the transgene product suggests that gene expression from AAVs may enhance Treg expansion in an IL-10 and KC-dependent manner (see Figure 3) [110].

10. Mechanisms and Evidence of rAAV Immunotoxicity

Complement activation has been implicated in several AAV-induced adverse events that occur 6–12 days after infection. Successful treatment of the majority of patients in previous studies with eculizumab, an antibody that inhibits the activation of the complement pathway, confirmed the involvement of this pathway in AAV-induced adverse effects [118,119]. Using a poorly understood mechanism, AAV doses of 5 × 1013 vg/kg or higher resulted in complement activation and thrombotic microangiopathies (TMA) in 9 out of 1400 patients in a spinal muscular atrophy (SMA) trial. The majority of those who suffered side effects received a higher dose of 1.1 × 1014 vg/kg [28,96,120]. TMA was also observed in 3 patients who received 1–3 × 1014 vg/kg and 1 patient who received 5 × 1013 vg/kg in two DMD trials that enrolled a total of 15 patients. Although the majority of patients in these trials were treated successfully, one patient in the SMA trial and one in the DMD trial died [24,28,121].
The prevalence of AAV-neutralizing antibodies (nAbs) varies, depending on the geographical location and serotype, with anti-AAV serotype 2 (AAV2) nAbs having the highest prevalence worldwide [122]. Cross-reactive antibodies that recognize several serotypes have also been identified. A longitudinal study of patients who received the AAV2 vector at varying doses for the treatment of hemophilia did not highlight serious safety concerns. However, a consistent increase in nAbs to AAV2, as well as AAV5 and AAV8, was noted [123]. Except for the involvement of AAV antibodies in complement activation-associated toxicities, information on the role of capsid-specific nAbs in toxicity is lacking. However, antibody responses to the transgene product have been shown to be toxic and seem to depend on the route of administration. Unlike intramuscular delivery, sub-retinal injection, portal vein administration, hepatic gene transfer, and injection into the central nervous system do not seem to induce transgene-specific antibody production and adverse effects [124,125,126,127].
Most AAV-mediated adverse effects are associated with cellular immunity. AAV capsid antigens are the most common inducers of aggressive adaptive immune responses. Both primary and secondary T cell responses have been implicated in hepatotoxicity following AAV infection. An increase in AAV-specific memory CD8+ T cells and diminished transgene expression were detected 4 weeks after infection with Factor IX-expressing AAVs in hemophilia B trials [128,129]. This delayed response suggests that the major contributor to hepatotoxicity is the slow primary T-cell responses driven by the capsid antigens from viral particle degradation. In the SMA trial, one-third of the patients showed an increase in CD8+ T cells and liver damage [28,130].

11. Overcoming AAV Immunotoxicity

11.1. Manipulating the Host’s Immune Response

Several studies have demonstrated the potential of general immunosuppressants in rAAV gene therapy [131,132]. A recent study demonstrated that antigen-specific tolerance can be induced by the co-administration of rAAVs with rapamycin. In this study, the co-administration of rapamycin, an inhibitor of the mammalian target of rapamycin (mTOR), with rAAVs resulted in capsid-specific effector T cell depletion and induced Treg proliferation [133]. The use of hydroxychloroquine, a drug used to treat malaria, to inhibit TLR9 activity an hour before AAV administration has also shown promising outcomes [134]. Complement inhibition with agents such as eculizumab and a synthetic peptide APL-9 has been successfully achieved to alleviate AAV-induced toxicity in clinical trials [29,118,119,135]. The administration of IgG-degrading enzymes before AAV injection may also avoid complement activation [136,137]. A recent study generated AAV-specific chimeric antigen receptor (CAR) Tregs. The injection of these CAR Tregs a week after the administration of a highly immunogenic AAV capsid suppressed effector T cell proliferation and cytotoxicity in mice (see Table 1) [138]. Although promising, general immune modulation is less favored in a clinical setting, as it may predispose patients to other infections.

11.2. AAV Genome Engineering

Supported by the superior efficacy of the FIX-Padua variant, highly efficient transgenes can be used to lower the AAV dose [139]. Lower hepatocyte-restricted transgene expression is key to tolerance. Moreover, trials using hepatocyte-specific promoters to express FIX and FVIII from AAVs showed a lack of transgene-specific immune responses and emphasized the need to use weaker liver-specific regulatory elements to drive transgene expression [140,141]. Using cis-acting regulatory modules that contain conserved clusters of transcription factor binding sites with high tissue-specific gene expression results in enhanced transgene expression [159]. A recent study reporting superior anti-HBV effects when using a liver-specific promoter to express saCas9 is encouraging (see Table 1) [86].
Recent studies have harnessed RNAi to design self-regulating AAVs and avoid transgene expression in specific cell types. Vectors can be made to carry a target of the tissue-specific miRNA or can be designed to co-express tissue-specific promoter-driven shRNA or pri-miRNA together with transgene mRNAs bearing the target site to the encoded shRNA/pri-miRNA [142]. This concept has been clearly demonstrated in CNS-targeting AAVs. Dorsal root ganglion (DRG) targeting by AAVs is a well-established contributor to AAV toxicity [160,161]. A recent study demonstrated that AAVs bearing dorsal root ganglion (DRG)-specific miR targets in the 3′ untranslated region of the transgene mRNA can mitigate AAV-associated DRG toxicity without affecting transduction in other parts of the brain [162]. A recent study showed that the AAV-mediated delivery of short immunomodulatory peptides, such as the hepatitis C virus non-structural protein 5A (NS5A), suppressed the unwanted activation of primary T-cells ex vivo. The expression of this proline-rich hydrophilic, 20-mer peptide represses the activation of memory T cell responses in the absence of CD3 (see Table 1) [147]. Designing anti-HBV AAVs that co-express this peptide may provide a strategy to overcome the anti-AAV adaptive immune response.
The depletion of CpG motifs in expression cassettes has been shown to diminish innate immune responses and enhance transgene expression [144,145]. The incorporation of TLR9-inhibiting non-coding DNA oligonucleotides into the AAV genome has previously been shown to significantly reduce innate immune and T cell response activation in mice and pigs. Although these results were promising, the vectors could only delay the onset of an inflammatory response in non-human primates, suggesting more complex anti-AAV innate immune responses in this model [163]. The double-stranded RNA molecules produced from the AAV ITR promoter trigger innate immune response-mediated toxicity [90,164,165]. Engineering the expression cassette, such that transcription from the 3′ITR is blocked, is a promising approach to avoid double-stranded RNA formation (see Table 1) [146].
As the only cis elements in rAAVs needed for viral replication, ITRs are the primary candidates for vector genome optimization to avoid immune-stimulatory effects. Although ITR modification is complicated by the inflexibility of the ITR sequences and their structure, the mutation of CpG motifs within the ITRs did not affect the therapeutic effects [143]. Interestingly, scAAV vectors induce higher transgene-specific CD8+ T and B cell responses than ssAAV [166]. This may be mediated by the double-stranded nature of the scAAV genome, which results in recognition by PRR, the activation of innate immune responses, and/or quicker transgene expression kinetics.

11.3. AAV Capsid Modification

The next generation of recombinant AAV vectors for gene therapy will benefit not only from emerging genome manipulation technologies but also from the improved ability to engineer capsids. Recent high-throughput single-cell analysis studies have unraveled the association between capsid tropism and the upregulation of specific genes, thereby establishing the genetic basis of AAV capsids’ promiscuous cell tropism [167]. Hence, capsid engineering to mediate tissue specificity may reduce the dose needed, de-target APCs, and evade nAbs. Thus far, several studies have demonstrated a range of modification possibilities while retaining capsid integrity, vector assembly, and high production yields. Bioengineering procedures to design capsids with higher hepatocyte transduction efficiencies, including rational designs to modify naturally occurring capsids and the design via directed evolution of novel cap genes, have been described and reviewed in the past [135,148,149,150,151].
The use of capsid engineering approaches to address the challenge of humoral immune responses and avoid complement activation has been investigated. These include using capsids with modified antibody epitopes and the production of novel capsid variants that can evade pre-existing antibodies via directed evolution [152,153,154]. The development of strategies to circumvent innate and cell-mediated immunity has recently gained momentum. A recent study showed the successful insertion of a TLR signal-inhibiting short peptide in the capsid without interfering with capsid assembly and vector yields. This capsid variant showed reduced Type 1 interferon and CD8+ T cell responses and delayed the formation of AAV-specific antibodies [155]. Phosphorylation and the subsequent ubiquitination of tyrosine residues on the AAV capsid enhance capsid degradation and antigen presentation. Hence, the mutation of the surface-exposed tyrosine residue can reduce T-cell activity and the associated toxicity [156,157]. The induction of tolerance by incorporating the MHC II epitopes derived from IgG in the AAV capsid also led to the proliferation of Tregs and the inhibition of CD8+ T cell activity (see Table 1) [158]. These techniques have provided limitless opportunities for capsid modification, aimed at overcoming the intrinsic challenges affecting the approval of AAV-based therapies. Although not yet leveraged, these strategies may benefit anti-HBV gene therapy designs.

12. Implications for AAV-Based Anti-HBV Gene Therapy Development

The demonstration that HBV infection enhances AAV transduction in an HBx-dependent manner supports the suitability of using AAVs for anti-HBV gene therapy [168]. AAVs are most likely to be safer and more effective for treatments targeting viruses that accumulate within immunosuppressive or immunotolerant environments, such as the liver. The demonstration that the simultaneous delivery of the same antigen to the liver and muscle can induce hepatic immune tolerance and mitigate the inflammatory responses induced by intramuscular gene transfer is an exciting observation [169,170]. The fact that both AAV and HBV can modulate cellular immunity to favor Treg responses may enhance the suppression of immunogenicity against AAV and boost the efficacy of AAVs in HBV-infected livers. AAV-induced IL-10 overproduction by KCs may suppress CD4+ T cell overactivity, as a result of HBV-induced MHC II overexpression by hepatocytes and favor CD4+ T cell differentiation into Tregs (see Figure 3). However, the correlation between elevated levels of Tregs with increased HBV loads in patients who are chronically infected with HBV is a concern [38,171]. Hence, harnessing the tolerogenic nature of the liver may be more complicated in patients with CHB. The treatment of patients with immunosuppressants before vector administration comes with the risk of making patients more susceptible to infection. However, inducing antigen-specific tolerance with immunosuppressants such as rapamycin may allow for repeat administrations of low-dose anti-HBV AAVs to achieve a therapeutic efficacy that is similar to that of one-off high-dose administration [133].
Lower transgene potency may result in the need for high AAV doses. The availability of highly potent anti-HBV sequences will enable the use of low vector doses, thus diminishing dose-dependent immune response-related adverse effects. This has been demonstrated by the promising outcomes reported in several pre-clinical studies using low-dose AAVs expressing anti-HBV gene editing nucleases or RNAi activators [16,22,85]. The superior anti-HBV effects reported when expressing saCas9 from a liver-specific promoter, the significant progress made in designing tools for the identification of CRISPR/Cas-mediated off-target effects, and limiting immunity against Cas 9 makes anti-HBV CRISPR/Cas therapies more attractive. AAV-mediated hemophilia gene therapy studies have shown that an immune response directed to the therapeutic gene product is linked to increased transgene product-mediated cytotoxic T-cell responses [172]. Evidence of the transgene products’ tolerance in the liver, induced by hepatocyte-restricted transgene expression, further supports the feasibility of AAV-mediated anti-HBV gene therapy [173,174]. Moreover, most of these anti-HBV AAV studies have used weaker liver-specific promoters to drive anti-HBV transgene expression. An example is the mouse transthyretin receptor (MTTR) promoter.
An efficient anti-HBV gene therapy vector must reach the viral reservoirs and persist in the liver. Hence, the availability of naturally existing AAV serotypes and artificially designed AAVs with high liver tropisms and the ability to form stable episomes is an attractive prospect. The resistance of some AAV serotypes to the humoral responses induced by community-acquired AAV infections may reduce complement activation [175]. AAV-based anti-HBV therapy designs may benefit from currently available capsid modification strategies. Capsid engineering to maximize hepatocyte transduction efficiency and de-target APCs is a promising approach. The insertion of TLR-inhibiting peptides and mutating tyrosine residues on the capsid surface to avoid the activation of TLR 9 signaling and antigen presentation, respectively, is another appealing strategy for anti-HBV gene therapy development. However, a recent demonstration that the AAV capsid assembly is stochastic and results in a highly heterogeneous capsid stoichiometry in one preparation is of concern. This may result in different liver transduction efficiencies between preparations and calls for careful consideration during capsid engineering [176].

13. Conclusions

Information on AAV immunogenicity has been instrumental in fast-tracking AAV-based gene therapy development. AAV genomes that carry transgene sequences with optimized potency against HBV, expressed from tissue-specific promoters, will reduce the vector doses required for clinical application. Anti-HBV gene therapy may benefit from adapting genome and capsid engineering strategies, developed for improving liver targeting, de-targeting from immune cells, and evading antibody and T-cell responses. However, this field of research has yet to take full advantage of these strategies. Information on how the nature of the immune response in HBV-infected tissues affects these efforts is required to inform the advancement of AAV-mediated anti-HBV gene therapy. Combining vector engineering and pharmacological interventions may be more effective in avoiding possible anti-HBV AAV toxicity. Whereas mouse studies have been instrumental in answering several key questions, they have also been misleading regarding some aspects of AAV immunogenicity. Developing better models of HBV infection is therefore crucial before the clinical translation of anti-HBV AAV therapies.

Author Contributions

R.J. wrote the sections on anti-AAV immunity and revised all drafts. M.D.D. wrote the section on capsid engineering and revised all drafts. N.M. wrote the section on HBV and AAV biology and revised all drafts. K.N. wrote the content on anti-HBV immunity and vaccine development. S.B. revised all the drafts. P.A. revised all the drafts and provided funding. M.B.M. conceptualized the idea, wrote the section on liver immunology, put the first draft together, revised all the drafts, finalized the manuscript for submission, and provided funding. All authors have read and agreed to the published version of the manuscript.

Funding

Financial support from the South African National Research Foundation (unique grant numbers: 118022 and 120383), the South African Medical Research Council of South Africa, the Female Academic Leadership Fellowship grant, the Carnegie Enabling grant, the Poliomyelitis Research Foundation, South African Research Chairs Initiative of the Department of Science and Technology (DST), The South African Research Chair in Cancer Biotechnology (UID 47904) is gratefully acknowledged.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Alberts, C.J.; Clifford, G.M.; Georges, D.; Negro, F.; A Lesi, O.; Hutin, Y.J.-F.; de Martel, C. Worldwide prevalence of hepatitis B virus and hepatitis C virus among patients with cirrhosis at country, region, and global levels: A systematic review. Lancet Gastroenterol. Hepatol. 2022, 7, e35576953. [Google Scholar] [CrossRef] [PubMed]
  2. Meireles, L.C.; Marinho, R.T.; Van Damme, P. Three decades of hepatitis B control with vaccination. World J. Hepatol. 2015, 7, 2127–2132. [Google Scholar] [CrossRef] [PubMed]
  3. Pattyn, J.; Hendrickx, G.; Vorsters, A.; Van Damme, P. Hepatitis B vaccines. J. Infect. Dis. 2021, 224 (Suppl. S2), S343–S351. [Google Scholar] [CrossRef] [PubMed]
  4. Wiedmann, M.; Liebert, U.G.; Oesen, U.; Porst, H.; Wiese, M.; Schroeder, S.; Halm, U.; Mössner, J.; Berr, F. Decreased immunogenicity of recombinant hepatitis B vaccine in chronic hepatitis C. Hepatology 2000, 31, 230–234. [Google Scholar] [CrossRef] [PubMed]
  5. Filippelli, M.; Lionetti, E.; Gennaro, A.; Lanzafame, A.; Arrigo, T.; Salpietro, C.; La Rosa, M.; Leonardi, S. Hepatitis B vaccine by intradermal route in non responder patients: An update. World J. Gastroenterol. 2014, 20, 10383–10394. [Google Scholar] [CrossRef]
  6. Meier, M.A.; Berger, C.T. A simple clinical score to identify likely hepatitis B vaccination non-responders—Data from a retrospective single center study. BMC Infect. Dis. 2020, 20, 891. [Google Scholar] [CrossRef]
  7. Körber, N.; Pohl, L.; Weinberger, B.; Grubeck-Loebenstein, B.; Wawer, A.; Knolle, P.A.; Roggendorf, H.; Protzer, U.; Bauer, T. Hepatitis 7. B Vaccine Non-Responders Show Higher Frequencies of CD24highCD38high Regulatory B Cells and Lower Levels of IL-10 Expression Compared to Responders. Front. Immunol. 2021, 12, 713351. [Google Scholar] [CrossRef]
  8. Saco, T.V.; Strauss, A.T.; Ledford, D.K. Hepatitis B vaccine nonresponders: Possible mechanisms and solutions. Ann. Allergy Asthma Immunol. 2018, 121, 320–327. [Google Scholar] [CrossRef]
  9. Karimkhanilouyi, S.; Ghorbian, S. Nucleic acid vaccines for hepatitis B and C virus. Infect. Genet. Evol. 2019, 75, 103968. [Google Scholar] [CrossRef]
  10. Zhao, H.; Shao, X.; Yu, Y.; Huang, L.; Amor, N.P.; Guo, K.; Weng, C.; Zhao, W.; Yang, A.; Hu, J.; et al. A therapeutic hepatitis B mRNA vaccine with strong immunogenicity and persistent virological suppression. BioRxiv 2022. [Google Scholar] [CrossRef]
  11. Zhang, X.; Wang, J.; Lu, J.; Li, R.; Zhao, S. Immunogenicity of adenovirus-vector vaccine targeting hepatitis B virus: Non-clinical safety assessment in non-human primates. Virol. J. 2018, 15, 111. [Google Scholar] [CrossRef]
  12. Nieto, K.; Salvetti, A. AAV Vectors Vaccines Against Infectious Diseases. Front. Immunol. 2014, 5, 5. [Google Scholar] [CrossRef]
  13. Travieso, T.; Li, J.; Mahesh, S.; Mello, J.D.F.R.E.; Blasi, M. The use of viral vectors in vaccine development. NPJ Vaccines 2022, 7, 75. [Google Scholar] [CrossRef]
  14. Martinez-Navio, J.M.; Fuchs, S.P.; Pantry, S.N.; Lauer, W.A.; Duggan, N.N.; Keele, B.F.; Rakasz, E.G.; Gao, G.; Lifson, J.D.; Desrosiers, R.C. Adeno-Associated Virus Delivery of Anti-HIV Monoclonal Antibodies Can Drive Long-Term Virologic Suppression. Immunity 2019, 50, 567–575.e5. [Google Scholar] [CrossRef]
  15. Tada, T.; Minnee, J.; Landau, N.R. Vectored immunoprophylaxis and treatment of SARS-CoV-2 infection in a preclinical model. Proc. Natl. Acad. Sci. USA 2023, 120, e2303509120. [Google Scholar] [CrossRef]
  16. Scott, T.; Moyo, B.; Nicholson, S.; Maepa, M.B.; Watashi, K.; Ely, A.; Weinberg, M.S.; Arbuthnot, P. ssAAVs containing cassettes encoding SaCas9 and guides targeting hepatitis B virus inactivate replication of the virus in cultured cells. Sci. Rep. 2017, 7, 7401. [Google Scholar] [CrossRef] [PubMed]
  17. Ely, A.; Naidoo, T.; Mufamadi, S.; Crowther, C.; Arbuthnot, P. Expressed anti-hbv primary microrna shuttles inhibit viral replication efficiently in vitro and in vivo. Mol. Ther. 2008, 16, 1105–1112. [Google Scholar] [CrossRef]
  18. Stone, D.; Long, K.R.; Loprieno, M.A.; Feelixge, H.S.D.S.; Kenkel, E.J.; Liley, R.M.; Rapp, S.; Roychoudhury, P.; Nguyen, T.; Stensland, L.; et al. CRISPR-Cas9 gene editing of hepatitis B virus in chronically infected humanized mice. Mol. Ther. Methods Clin. Dev. 2021, 20, 258–275. [Google Scholar] [CrossRef] [PubMed]
  19. Kayesh, M.E.H.; Amako, Y.; Hashem, A.; Murakami, S.; Ogawa, S.; Yamamoto, N.; Hifumi, T.; Miyoshi, N.; Sugiyama, M.; Tanaka, Y.; et al. Development of an in vivo delivery system for CRISPR/Cas9-mediated targeting of hepatitis B virus cccDNA. Virus Res. 2020, 290, 198191. [Google Scholar] [CrossRef] [PubMed]
  20. Marimani, M.D.; Ely, A.; Buff, M.C.; Bernhardt, S.; Engels, J.W.; Scherman, D.; Escriou, V.; Arbuthnot, P. Inhibition of replication of hepatitis B virus in transgenic mice following administration of hepatotropic lipoplexes containing guanidinopropyl-modified siRNAs. J. Control. Release 2015, 209, 198–206. [Google Scholar] [CrossRef]
  21. Chen, B.; Chen, Y.; Li, J.; Wang, C.; Song, W.; Wen, Y.; Ying, T. A Single Dose of Anti-HBsAg Antibody-Encoding mRNA-LNPs Suppressed HBsAg Expression: A Potential Cure of Chronic Hepatitis B Virus Infection. MBio 2022, 13, e01612–e01622. [Google Scholar] [CrossRef] [PubMed]
  22. Maepa, M.B.; Ely, A.; Grayson, W.; Arbuthnot, P. Sustained Inhibition of HBV Replication In Vivo after Systemic Injection of AAVs Encoding Artificial Antiviral Primary MicroRNAs. Mol. Ther. Nucleic Acids 2017, 7, 190–199. [Google Scholar] [CrossRef] [PubMed]
  23. Au, H.K.E.; Isalan, M.; Mielcarek, M. Gene Therapy Advances: A Meta-Analysis of AAV Usage in Clinical Settings. Front. Med. 2021, 8, 809118. [Google Scholar] [CrossRef]
  24. Shen, W.; Liu, S.; Ou, L. rAAV immunogenicity, toxicity, and durability in 255 clinical trials: A meta-analysis. Front. Immunol. 2022, 13, 1001263. [Google Scholar] [CrossRef]
  25. Philippidis, A. Fourth Boy Dies in Clinical Trial of Astellas’ AT132. Hum. Gene Ther. 2021, 32, 1008–1010. [Google Scholar] [CrossRef] [PubMed]
  26. Philippidis, A. Novartis Confirms Deaths of Two Patients Treated with Gene Therapy Zolgensma. Hum. Gene Ther. 2022, 33, 842–844. [Google Scholar] [CrossRef] [PubMed]
  27. Lek, A.; Wong, B.; Keeler, A.; Blackwood, M.; Ma, K.; Huang, S.; Sylvia, K.; Batista, A.R.; Artinian, R.; Kokoski, D.; et al. Death after high-dose rAAV9 gene therapy in a patient with duchenne’s muscular dystrophy. N. Engl. J. Med. 2023, 389, 1203–1210. [Google Scholar] [CrossRef]
  28. Ertl, H.C.J. Immunogenicity and toxicity of AAV gene therapy. Front. Immunol. 2022, 13, 975803. [Google Scholar] [CrossRef]
  29. Arjomandnejad, M.; Dasgupta, I.; Flotte, T.R.; Keeler, A.M. Immunogenicity of Recombinant Adeno-Associated Virus (AAV) Vectors for Gene Transfer. BioDrugs 2023, 37, 311–329. [Google Scholar] [CrossRef]
  30. Kramvis, A. Genotypes and genetic variability of hepatitis B virus. Intervirology 2014, 57, 141–150. [Google Scholar] [CrossRef]
  31. Tsukuda, S.; Watashi, K. Hepatitis B virus biology and life cycle. Antivir. Res. 2020, 182, 104925. [Google Scholar] [CrossRef] [PubMed]
  32. Tan, A.; Koh, S.; Bertoletti, A. Immune Response in Hepatitis B Virus Infection. Cold Spring Harb. Perspect. Med. 2015, 5, a021428. [Google Scholar] [CrossRef] [PubMed]
  33. Chisari, F.V. Cytotoxic T cells and viral hepatitis. J. Clin. Investig. 1997, 99, 1472–1477. [Google Scholar] [CrossRef] [PubMed]
  34. Bertoletti, A.; Ferrari, C. Adaptive immunity in HBV infection. J. Hepatol. 2016, 64 (Suppl. S1), S71–S83. [Google Scholar] [CrossRef] [PubMed]
  35. Ye, B.; Liu, X.; Li, X.; Kong, H.; Tian, L.; Chen, Y. T-cell exhaustion in chronic hepatitis B infection: Current knowledge and clinical significance. Cell Death Dis. 2015, 6, e1694. [Google Scholar] [CrossRef] [PubMed]
  36. Lin, N.; Yin, W.; Miller, H.; Byazrova, M.G.; Herrada, A.A.; Benlagha, K.; Lee, P.; Guan, F.; Lei, J.; Gong, Q.; et al. The role of regulatory T cells and follicular T helper cells in HBV infection. Front. Immunol. 2023, 14, 1169601. [Google Scholar] [CrossRef] [PubMed]
  37. Jung, M.K.; Shin, E.-C. Regulatory T Cells in Hepatitis B and C Virus Infections. Immune Netw. 2016, 16, 330–336. [Google Scholar] [CrossRef]
  38. Stoop, J.N.; van der Molen, R.G.; Baan, C.C.; van der Laan, L.J.W.; Kuipers, E.J.; Kusters, J.G.; Janssen, H.L.A. Regulatory T cells contribute to the impaired immune response in patients with chronic hepatitis B virus infection. Hepatology 2005, 41, 771–778. [Google Scholar] [CrossRef]
  39. Yang, G.; Liu, A.; Xie, Q.; Guo, T.B.; Wan, B.; Zhou, B.; Zhang, J.Z. Association of CD4+CD25+Foxp3+ regulatory T cells with chronic activity and viral clearance in patients with hepatitis B. Int. Immunol. 2007, 19, 133–140. [Google Scholar] [CrossRef]
  40. Tang, R.; Lei, Z.; Wang, X.; Qi, Q.; He, J.; Liu, D.; Wang, X.; Chen, X.; Zhu, J.; Li, Y.; et al. Hepatitis B envelope antigen increases Tregs by converting CD4+CD25- T cells into CD4+CD25+Foxp3+ Tregs. Exp. Ther. Med. 2020, 20, 3679–3686. [Google Scholar] [CrossRef]
  41. Kondo, Y.; Ueno, Y.; Kobayashi, K.; Kakazu, E.; Shiina, M.; Inoue, J.; Tamai, K.; Wakui, Y.; Tanaka, Y.; Ninomiya, M.; et al. Hepatitis B virus replication could enhance regulatory T cell activity by producing soluble heat shock protein 60 from hepatocytes. J. Infect. Dis. 2010, 202, 202–213. [Google Scholar] [CrossRef]
  42. Qiu, H.; Wang, N.; Lin, D.; Yuan, Y.; Li, J.; Mao, D.; Meng, Y. The positive feedback loop of furin and TGFβ1 enhances the immune responses of Tregs to hepatocellular carcinoma cells and hepatitis B virus in vitro. Cell Biol. Int. 2022, 46, 1215–1226. [Google Scholar] [CrossRef] [PubMed]
  43. Zhang, C.; Gao, Y.; Du, C.; Markowitz, G.J.; Fu, J.; Zhang, Z.; Liu, C.; Qin, W.; Wang, H.; Wang, F.; et al. Hepatitis B–Induced IL8 Promotes Hepatocellular Carcinoma Venous Metastasis and Intrahepatic Treg Accumulation. Cancer Res 2021, 81, 2386–2398. [Google Scholar] [CrossRef] [PubMed]
  44. Cai, Y.; Yin, W. The Multiple Functions of B Cells in Chronic HBV Infection. Front. Immunol. 2020, 11, 582292. [Google Scholar] [CrossRef] [PubMed]
  45. Ely, A.; Naidoo, T.; Arbuthnot, P. Efficient silencing of gene expression with modular trimeric Pol II expression cassettes comprising microRNA shuttles. Nucleic Acids Res. 2009, 37, e91. [Google Scholar] [CrossRef] [PubMed]
  46. Giering, J.C.; Grimm, D.; Storm, T.A.; Kay, M.A. Expression of shRNA from a tissue-specific pol ii promoter is an effective and safe RNAi therapeutic. Mol. Ther. 2008, 16, 1630. [Google Scholar] [CrossRef]
  47. Jacobs, R.; Singh, P.; Smith, T.; Arbuthnot, P.; Maepa, M.B. Prospects of viral vector-mediated delivery of sequences encoding anti-HBV designer endonucleases. Gene Ther. 2022, 1–8. [Google Scholar] [CrossRef]
  48. Kim, H.; Kim, J.-S. A guide to genome engineering with programmable nucleases. Nat. Rev. Genet. 2014, 15, 321–334. [Google Scholar] [CrossRef]
  49. Cradick, T.J.; Keck, K.; Bradshaw, S.; Jamieson, A.C.; McCaffrey, A.P. Zinc-finger nucleases as a novel therapeutic strategy for targeting hepatitis B virus DNAs. Mol. Ther. 2010, 18, 947–954. [Google Scholar] [CrossRef]
  50. Weber, N.D.; Stone, D.; Sedlak, R.H.; Feelixge, H.S.D.S.; Roychoudhury, P.; Schiffer, J.T.; Aubert, M.; Jerome, K.R. AAV-mediated delivery of zinc finger nucleases targeting hepatitis B virus inhibits active replication. PLoS ONE 2014, 9, e97579. [Google Scholar] [CrossRef]
  51. Bloom, K.; Ely, A.; Mussolino, C.; Cathomen, T.; Arbuthnot, P. Inactivation of hepatitis B virus replication in cultured cells and in vivo with engineered transcription activator-like effector nucleases. Mol. Ther. 2013, 21, 1889–1897. [Google Scholar] [CrossRef]
  52. Chen, J.; Zhang, W.; Lin, J.; Wang, F.; Wu, M.; Chen, C.; Zheng, Y.; Peng, X.; Li, J.; Yuan, Z. An efficient antiviral strategy for targeting hepatitis B Virus genome using transcription activator-like effector nucleases. Mol. Ther. 2014, 22, 303–311. [Google Scholar] [CrossRef] [PubMed]
  53. Smith, T.; Singh, P.; Chmielewski, K.O.; Bloom, K.; Cathomen, T.; Arbuthnot, P.; Ely, A. Improved Specificity and Safety of Anti-Hepatitis B Virus TALENs Using Obligate Heterodimeric FokI Nuclease Domains. Viruses 2021, 13, 1344. [Google Scholar] [CrossRef] [PubMed]
  54. Mali, P.; Yang, L.; Esvelt, K.M.; Aach, J.; Guell, M.; DiCarlo, J.E.; Norville, J.E.; Church, G.M. RNA-guided human genome engineering via Cas9. Science 2013, 339, 823–826. [Google Scholar] [CrossRef] [PubMed]
  55. Marraffini, L.A. The CRISPR-Cas system of Streptococcus pyogenes: Function and applications. In Streptococcus pyogenes: Basic Biology to Clinical Manifestations; University of Oklahoma Health Sciences Center: Oklahoma City, OK, USA, 2016. [Google Scholar]
  56. Yourik, P.; Fuchs, R.T.; Mabuchi, M.; Curcuru, J.L.; Robb, G.B. Staphylococcus aureus Cas9 is a multiple-turnover enzyme. RNA 2019, 25, 35–44. [Google Scholar] [CrossRef] [PubMed]
  57. Seeger, C.; Sohn, J.A. Targeting hepatitis B virus with CRISPR/Cas9. Mol. Ther. Nucleic Acids 2014, 3, e216. [Google Scholar] [CrossRef] [PubMed]
  58. Dong, C.; Qu, L.; Wang, H.; Wei, L.; Dong, Y.; Xiong, S. Targeting hepatitis B virus cccDNA by CRISPR/Cas9 nuclease efficiently inhibits viral replication. Antivir. Res. 2015, 118, 110–117. [Google Scholar] [CrossRef] [PubMed]
  59. Wang, J. Dual gRNAs guided CRISPR/Cas9 system inhibits hepatitis B virus replication. World J. Gastroenterol. 2015, 21, 9554–9565. [Google Scholar] [CrossRef]
  60. Ramanan, V.; Shlomai, A.; Cox, D.B.; Schwartz, R.E.; Michailidis, E.; Bhatta, A.; Scott, D.A.; Zhang, F.; Rice, C.M.; Bhatia, S.N. CRISPR/Cas9 cleavage of viral DNA efficiently suppresses hepatitis B virus. Sci. Rep. 2015, 5, srep10833. [Google Scholar] [CrossRef]
  61. Zhen, S.; Hua, L.; Liu, Y.H.; Gao, L.C.; Fu, J.; Wan, D.Y.; Gao, X. Harnessing the clustered regularly interspaced short palindromic repeat (CRISPR)/CRISPR-associated Cas9 system to disrupt the hepatitis B virus. Gene Ther. 2015, 22, 404–412. [Google Scholar] [CrossRef]
  62. Kennedy, E.M.; Bassit, L.C.; Mueller, H.; Kornepati, A.V.; Bogerd, H.P.; Nie, T.; Chatterjee, P.; Javanbakht, H.; Schinazi, R.F.; Cullen, B.R. Suppression of hepatitis B virus DNA accumulation in chronically infected cells using a bacterial CRISPR/Cas RNA-guided DNA endonuclease. Virology 2015, 476, 196–205. [Google Scholar] [CrossRef] [PubMed]
  63. Li, H.; Sheng, C.; Wang, S.; Yang, L.; Liang, Y.; Huang, Y.; Liu, H.; Li, P.; Yang, C.; Yang, X.; et al. Removal of integrated hepatitis B virus DNA using CRISPR-Cas9. Front. Cell. Infect. Microbiol. 2017, 7, 91. [Google Scholar] [CrossRef] [PubMed]
  64. Lin, S.-R.; Yang, H.-C.; Kuo, Y.-T.; Liu, C.-J.; Yang, T.-Y.; Sung, K.-C.; Lin, Y.-Y.; Wang, H.-Y.; Wang, C.-C.; Shen, Y.-C.; et al. The CRISPR/Cas9 system facilitates clearance of the intrahepatic HBV templates in vivo. Mol. Ther. Nucleic Acids 2014, 3, e186. [Google Scholar] [CrossRef] [PubMed]
  65. Fu, Y.; Foden, J.A.; Khayter, C.; Maeder, M.L.; Reyon, D.; Joung, J.K.; Sander, J.D. High-frequency off-target mutagenesis induced by CRISPR-Cas nucleases in human cells. Nat. Biotechnol. 2013, 31, 822–826. [Google Scholar] [CrossRef] [PubMed]
  66. Pattanayak, V.; Lin, S.; Guilinger, J.P.; Ma, E.; Doudna, J.A.; Liu, D.R. High-throughput profiling of off-target DNA cleavage reveals RNA-programmed Cas9 nuclease specificity. Nat. Biotechnol. 2013, 31, 839–843. [Google Scholar] [CrossRef] [PubMed]
  67. Kang, S.-H.; Lee, W.-J.; An, J.-H.; Lee, J.-H.; Kim, Y.-H.; Kim, H.; Oh, Y.; Park, Y.-H.; Jin, Y.B.; Jun, B.-H.; et al. Prediction-based highly sensitive CRISPR off-target validation using target-specific DNA enrichment. Nat. Commun. 2020, 11, 3596. [Google Scholar] [CrossRef] [PubMed]
  68. Ran, F.A.; Hsu, P.D.; Lin, C.-Y.; Gootenberg, J.S.; Konermann, S.; Trevino, A.E.; Scott, D.A.; Inoue, A.; Matoba, S.; Zhang, Y.; et al. Double Nicking by RNA-Guided CRISPR Cas9 for enhanced genome editing specificity. Cell 2013, 154, 1380–1389. [Google Scholar] [CrossRef]
  69. Karimova, M.; Beschorner, N.; Dammermann, W.; Chemnitz, J.; Indenbirken, D.; Bockmann, J.-H.; Grundhoff, A.; Lüth, S.; Buchholz, F.; Wiesch, J.S.Z.; et al. CRISPR/Cas9 nickase-mediated disruption of hepatitis B virus open reading frame S and X. Sci. Rep. 2015, 5, 13734. [Google Scholar] [CrossRef]
  70. Ewaisha, R.; Anderson, K.S. Immunogenicity of CRISPR therapeutics—Critical considerations for clinical translation. Front. Bioeng. Biotechnol. 2023, 11, 1138596. [Google Scholar] [CrossRef]
  71. Samulski, R.J.; Muzyczka, N. AAV-Mediated Gene Therapy for Research and Therapeutic Purposes. Annu. Rev. Virol. 2014, 1, 427–451. [Google Scholar] [CrossRef]
  72. Xie, Q.; Bu, W.; Bhatia, S.; Hare, J.; Somasundaram, T.; Azzi, A.; Chapman, M.S. The atomic structure of adeno-associated virus (AAV-2), a vector for human gene therapy. Proc. Natl. Acad. Sci. USA 2002, 99, 10405–10410. [Google Scholar] [CrossRef] [PubMed]
  73. Maclachlan, N.J.; Dubovi, E.J. (Eds.) Chapter 12—Parvoviridae. In Fenner’s Veterinary Virology, 5th ed.; Academic Press: Boston, MA, USA, 2017; pp. 245–257. [Google Scholar]
  74. Srivastava, A. In vivo tissue-tropism of adeno-associated viral vectors. Curr. Opin. Virol. 2016, 21, 75–80. [Google Scholar] [CrossRef] [PubMed]
  75. Issa, S.S.; Shaimardanova, A.A.; Solovyeva, V.V.; Rizvanov, A.A. Various AAV Serotypes and Their Applications in Gene Therapy: An Overview. Cells 2023, 12, 785. [Google Scholar] [CrossRef]
  76. Sabatino, D.E.; Bushman, F.D.; Chandler, R.J.; Crystal, R.G.; Davidson, B.L.; Dolmetsch, R.; Eggan, K.C.; Gao, G.; Gil-Farina, I.; Kay, M.A.; et al. Evaluating the state of the science for adeno-associated virus integration: An integrated perspective. Mol. Ther. 2022, 30, 2646–2663. [Google Scholar] [CrossRef] [PubMed]
  77. Meyer, N.L.; Chapman, M.S. Adeno-associated virus (AAV) cell entry: Structural insights. Trends Microbiol. 2022, 30, 432–451. [Google Scholar] [CrossRef] [PubMed]
  78. Yang, J.; Zhou, W.; Zhang, Y.; Zidon, T.; Ritchie, T.; Engelhardt, J.F. Concatamerization of adeno-associated virus circular genomes occurs through intermolecular recombination. J. Virol. 1999, 73, 9468–9477. [Google Scholar] [CrossRef] [PubMed]
  79. McCarty, D.M.; Fu, H.; Monahan, P.E.; Toulson, C.E.; Naik, P.; Samulski, R.J. Adeno-associated virus terminal repeat (TR) mutant generates self-complementary vectors to overcome the rate-limiting step to transduction in vivo. Gene Ther. 2003, 10, 2112–2118. [Google Scholar] [CrossRef]
  80. Ferrari, F.K.; Samulski, T.; Shenk, T.; Samulski, R.J. Second-strand synthesis is a rate-limiting step for efficient transduction by recombinant adeno-associated virus vectors. J. Virol. 1996, 70, 3227–3234. [Google Scholar] [CrossRef]
  81. McCarty, D.M. Self-complementary AAV Vectors; Advances and Applications. Mol. Ther. 2008, 16, 1648–1656. [Google Scholar] [CrossRef]
  82. Ye, L.; Kan, F.; Yan, T.; Cao, J.; Zhang, L.; Wu, Z.; Li, W. Enhanced antiviral and antifibrotic effects of short hairpin RNAs targeting HBV and TGF-β in HBV-persistent mice. Sci. Rep. 2017, 7, 3860. [Google Scholar] [CrossRef] [PubMed]
  83. Chen, C.C.; Ko, T.-M.; Ma, H.-I.; Wu, H.-L.; Xiao, X.; Li, J.; Chang, C.-M.; Wu, P.-Y.; Chen, C.-H.; Han, J.-M.; et al. Long-term inhibition of hepatitis B virus in transgenic mice by double-stranded adeno-associated virus 8-delivered short hairpin RNA. Gene Ther. 2007, 14, 11–19. [Google Scholar] [CrossRef] [PubMed]
  84. Liu, Y.; Zhao, M.; Gong, M.; Xu, Y.; Xie, C.; Deng, H.; Li, X.; Wu, H.; Wang, Z. Inhibition of hepatitis B virus replication via HBV DNA cleavage by Cas9 from Staphylococcus aureus. Antivir. Res. 2018, 152, 58–67. [Google Scholar] [CrossRef] [PubMed]
  85. Li, H.; Sheng, C.; Liu, H.; Wang, S.; Zhao, J.; Yang, L.; Song, H. Inhibition of HBV expression in HBV transgenic mice using AAV-delivered CRISPR-SaCas9. Front. Immunol. 2018, 9, 2080. [Google Scholar] [CrossRef] [PubMed]
  86. Yan, K.; Feng, J.; Liu, X.; Wang, H.; Li, Q.; Li, J.; Xu, T.; Sajid, M.; Ullah, H.; Zhou, L.; et al. Inhibition of Hepatitis B Virus by AAV8-Derived CRISPR/SaCas9 Expressed from Liver-Specific Promoters. Front. Microbiol. 2021, 12, 665184. [Google Scholar] [CrossRef] [PubMed]
  87. 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. [Google Scholar] [CrossRef]
  88. Shitik, E.M.; Shalik, I.K.; Yudkin, D.V. AAV- based vector improvements unrelated to capsid protein modification. Front. Med. 2023, 10, 1106085. [Google Scholar] [CrossRef] [PubMed]
  89. Logan, G.J.; Dane, A.P.; Hallwirth, C.V.; Smyth, C.M.; Wilkie, E.E.; Amaya, A.K.; Zhu, E.; Khandekar, N.; Ginn, S.L.; Liao, S.H.Y.; et al. Identification of liver-specific enhancer–promoter activity in the 3′ untranslated region of the wild-type AAV2 genome. Nat. Genet. 2017, 49, 1267–1273. [Google Scholar] [CrossRef]
  90. Keiser, M.S.; Ranum, P.T.; Yrigollen, C.M.; Carrell, E.M.; Smith, G.R.; Muehlmatt, A.L.; Chen, Y.H.; Stein, J.M.; Wolf, R.L.; Radaelli, E.; et al. Toxicity after AAV delivery of RNAi expression constructs into nonhuman primate brain. Nat. Med. 2021, 27, 1982–1989. [Google Scholar] [CrossRef]
  91. Dalwadi, D.A.; Torrens, L.; Abril-Fornaguera, J.; Pinyol, R.; Willoughby, C.; Posey, J.; Llovet, J.M.; Lanciault, C.; Russell, D.W.; Grompe, M.; et al. Liver Injury Increases the Incidence of HCC following AAV Gene Therapy in Mice. Mol. Ther. 2021, 29, 680–690. [Google Scholar] [CrossRef]
  92. Chandler, R.J.; Sands, M.S.; Venditti, C.P. Recombinant Adeno-Associated Viral Integration and Genotoxicity: Insights from Animal Models. Hum. Gene Ther. 2017, 28, 314–322. [Google Scholar] [CrossRef]
  93. Ferla, R.; Alliegro, M.; Dell’anno, M.; Nusco, E.; Cullen, J.M.; Smith, S.N.; Wolfsberg, T.G.; O’donnell, P.; Wang, P.; Nguyen, A.-D.; et al. Low incidence of hepatocellular carcinoma in mice and cats treated with systemic adeno-associated viral vectors. Mol. Ther. Methods Clin. Dev. 2021, 20, 247–257. [Google Scholar] [CrossRef]
  94. Schäffer, A.A.; Dominguez, D.A.; Chapman, L.M.; Gertz, E.M.; Budhu, A.; Forgues, M.; Chaisaingmongkol, J.; Rabibhadana, S.; Pupacdi, B.; Wu, X.; et al. Integration of adeno-associated virus (AAV) into the genomes of most Thai and Mongolian liver cancer patients does not induce oncogenesis. BMC Genom. 2021, 22, 814. [Google Scholar] [CrossRef]
  95. Ertl, H.C. Preclinical models to assess the immunogenicity of AAV vectors. Cell. Immunol. 2019, 342, 103722. [Google Scholar] [CrossRef] [PubMed]
  96. Guillou, J.; de Pellegars, A.; Porcheret, F.; Frémeaux-Bacchi, V.; Allain-Launay, E.; Debord, C.; Denis, M.; Péréon, Y.; Barnérias, C.; Desguerre, I.; et al. Fatal thrombotic microangiopathy case following adeno-associated viral SMN gene therapy. Blood Adv. 2022, 6, 4266–4270. [Google Scholar] [CrossRef] [PubMed]
  97. Noris, M.; Remuzzi, G. Overview of complement activation and regulation. Semin. Nephrol. 2013, 33, 479–492. [Google Scholar] [CrossRef] [PubMed]
  98. Smith, C.J.; Ross, N.; Kamal, A.; Kim, K.Y.; Kropf, E.; Deschatelets, P.; Francois, C.; Quinn, W.J.; Singh, I.; Majowicz, A.; et al. Pre-existing humoral immunity and complement pathway contribute to immunogenicity of adeno-associated virus (AAV) vector in human blood. Front. Immunol. 2022, 13, 999021. [Google Scholar] [CrossRef] [PubMed]
  99. West, C.; Federspiel, J.D.; Rogers, K.; Khatri, A.; Rao-Dayton, S.; Ocana, M.F.; Lim, S.; D’Antona, A.M.; Casinghino, S.; Somanathan, S. Complement Activation by Adeno-Associated Virus-Neutralizing Antibody Complexes. Hum. Gene Ther. 2023, 34, 554–566. [Google Scholar] [CrossRef] [PubMed]
  100. Fong, S.; Handyside, B.; Sihn, C.-R.; Liu, S.; Zhang, L.; Xie, L.; Murphy, R.; Galicia, N.; Yates, B.; Minto, W.C.; et al. Induction of ER Stress by an AAV5 BDD FVIII Construct Is Dependent on the Strength of the Hepatic-Specific Promoter. Mol. Ther. Methods Clin. Dev. 2020, 18, 620–630. [Google Scholar] [CrossRef] [PubMed]
  101. Johnson, J.S.; Gentzsch, M.; Zhang, L.; Ribeiro, C.M.P.; Kantor, B.; Kafri, T.; Pickles, R.J.; Samulski, R.J. AAV exploits subcellular stress associated with inflammation, endoplasmic reticulum expansion, and misfolded proteins in models of cystic fibrosis. PLoS Pathog. 2011, 7, e1002053. [Google Scholar] [CrossRef] [PubMed]
  102. Martino, A.T.; Markusic, D.M. Immune Response Mechanisms against AAV Vectors in Animal Models. Mol. Ther. Methods Clin. Dev. 2020, 17, 198–208. [Google Scholar] [CrossRef]
  103. Nidetz, N.F.; McGee, M.C.; Tse, L.V.; Li, C.; Cong, L.; Li, Y.; Huang, W. Adeno-associated viral vector-mediated immune responses: Understanding barriers to gene delivery. Pharmacol. Ther. 2020, 207, 107453. [Google Scholar] [CrossRef]
  104. Shirley, J.L.; de Jong, Y.P.; Terhorst, C.; Herzog, R.W. Immune Responses to Viral Gene Therapy Vectors. Mol. Ther. 2020, 28, 709–722. [Google Scholar] [CrossRef]
  105. Sakaguchi, S.; Yamaguchi, T.; Nomura, T.; Ono, M. Regulatory T cells and immune tolerance. Cell 2008, 133, 775–787. [Google Scholar] [CrossRef] [PubMed]
  106. Herzog, R.W.; Cooper, M.; Perrin, G.Q.; Biswas, M.; Martino, A.T.; Morel, L.; Terhorst, C.; Hoffman, B.E. Regulatory T cells and TLR9 activation shape antibody formation to a secreted transgene product in AAV muscle gene transfer. Cell. Immunol. 2019, 342, 103682. [Google Scholar] [CrossRef] [PubMed]
  107. Wang, S.; Xu, H.; Wang, Y.; Ma, J.; Mao, C.; Shao, Q.; Ma, B.; Xu, W.; Yang, S. Regulatory T cells induced by rAAV carrying the forkhead box P3 gene prevent autoimmune thyroiditis in mice. Int. J. Mol. Med. 2006, 18, 1193–1199. [Google Scholar] [CrossRef]
  108. Thomson, A.W.; Knolle, P.A. Antigen-presenting cell function in the tolerogenic liver environment. Nat. Rev. Immunol. 2010, 10, 753–766. [Google Scholar] [CrossRef] [PubMed]
  109. Keeler, G.D.; Markusic, D.M.; Hoffman, B.E. Liver induced transgene tolerance with AAV vectors. Cell. Immunol. 2019, 342, 103728. [Google Scholar] [CrossRef] [PubMed]
  110. 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. [Google Scholar] [CrossRef]
  111. Heymann, F.; Peusquens, J.; Ludwig-Portugall, I.; Kohlhepp, M.; Ergen, C.; Niemietz, P.; Martin, C.; van Rooijen, N.; Ochando, J.C.; Randolph, G.J.; et al. Liver inflammation abrogates immunological tolerance induced by Kupffer cells. Hepatology 2015, 62, 279–291. [Google Scholar] [CrossRef]
  112. Warren, A.; Le Couteur, D.G.; Fraser, R.; Bowen, D.G.; McCaughan, G.W.; Bertolino, P. T lymphocytes interact with hepatocytes through fenestrations in murine liver sinusoidal endothelial cells. Hepatology 2006, 44, 1182–1190. [Google Scholar] [CrossRef]
  113. Herkel, J.; Jagemann, B.; Wiegard, C.; Lazaro, J.F.G.; Lueth, S.; Kanzler, S.; Blessing, M.; Schmitt, E.; Lohse, A.W. MHC class II-expressing hepatocytes function as antigen-presenting cells and activate specific CD4 T lymphocyutes. Hepatology 2003, 37, 1079–1085. [Google Scholar] [CrossRef] [PubMed]
  114. Burghardt, S.; Claass, B.; Erhardt, A.; Karimi, K.; Tiegs, G. Hepatocytes induce Foxp3+ regulatory T cells by Notch signaling. J. Leukoc. Biol. 2014, 96, 571–577. [Google Scholar] [CrossRef] [PubMed]
  115. Hösel, M.; Broxtermann, M.; Janicki, H.; Esser, K.; Arzberger, S.; Hartmann, P.; Gillen, S.; Kleeff, J.; Stabenow, D.; Odenthal, M.; et al. Toll-like receptor 2-mediated innate immune response in human nonparenchymal liver cells toward adeno-associated viral vectors. Hepatology 2012, 55, 287–297. [Google Scholar] [CrossRef] [PubMed]
  116. 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 in vivo viral gene transfer. Blood 2004, 104, 969–977. [Google Scholar] [CrossRef] [PubMed]
  117. Cao, O.; Dobrzynski, E.; Wang, L.; Nayak, S.; Mingle, B.; Terhorst, C.; Herzog, R.W. Induction and role of regulatory CD4+CD25+ T cells in tolerance to the transgene product following hepatic in vivo gene transfer. Blood 2007, 110, 1132–1140. [Google Scholar] [CrossRef] [PubMed]
  118. Witte, D.; Hartmann, H.; Drube, J.; Haffner, D.; Illsinger, S. Thrombotic Microangiopathy (TMA) after Gene Replacemant Therapy (GRT) Due to Spinal Muscular Atrophy: Case Summary and Recommendations for Treatment. Klin. Padiatr. 2022, 234, 42–47. [Google Scholar]
  119. Yazaki, K.; Sakuma, S.; Hikita, N.; Fujimaru, R.; Hamazaki, T. Child Neurology: Pathologically Confirmed Thrombotic Microangiopathy Caused by Onasemnogene Abeparvovec Treatment for SMA. Neurology 2022, 98, 808–813. [Google Scholar] [CrossRef]
  120. Chand, D.H.; Zaidman, C.; Arya, K.; Millner, R.; Farrar, M.A.; Mackie, F.E.; Goedeker, N.L.; Dharnidharka, V.R.; Dandamudi, R.; Reyna, S.P. Thrombotic Microangiopathy Following Onasemnogene Abeparvovec for Spinal Muscular Atrophy: A Case Series. J. Pediatr. 2021, 231, 265–268. [Google Scholar] [CrossRef]
  121. Kishimoto, T.K.; Samulski, R.J. Addressing high dose AAV toxicity–‘one and done’or ‘slower and lower’? Expert Opin. Biol. Ther. 2022, 22, 1067–1071. [Google Scholar] [CrossRef]
  122. 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. [Google Scholar] [CrossRef]
  123. George, L.A.; Ragni, M.V.; Rasko, J.E.; Raffini, L.J.; Samelson-Jones, B.J.; Ozelo, M.; Hazbon, M.; Runowski, A.R.; Wellman, J.A.; Wachtel, K.; et al. Long-term follow-up of the first in human intravascular delivery of AAV for gene transfer: AAV2-hFIX16 for severe hemophilia B. Mol. Ther. 2020, 28, 2073–2082. [Google Scholar] [CrossRef] [PubMed]
  124. Bainbridge, J.W.; Smith, A.J.; Barker, S.S.; Robbie, S.; Henderson, R.; Balaggan, K.; Ali, R.R. Effect of gene therapy on visual function in Leber’s congenital amaurosis. N. Engl. J. Med. 2008, 358, 2231–2239. [Google Scholar] [CrossRef] [PubMed]
  125. Sherman, A.; Schlachterman, A.; Cooper, M.; Merricks, E.P.; Raymer, R.A.; Bellinger, D.A.; Nichols, T.C. Portal vein delivery of viral vectors for gene therapy for hemophilia. Gene Correct. Methods Protoc. 2014, 1114, 413–426. [Google Scholar]
  126. Samaranch, L.; Salegio, E.A.; Sebastian, W.S.; Kells, A.P.; Foust, K.D.; Bringas, J.R.; Lamarre, C.; Forsayeth, J.; Kaspar, B.K.; Bankiewicz, K.S.; et al. Adeno-associated virus serotype 9 transduction in the central nervous system of nonhuman primates. Hum. Gene Ther. 2012, 23, 382–389. [Google Scholar] [CrossRef] [PubMed]
  127. 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. [Google Scholar] [CrossRef] [PubMed]
  128. Manno, C.S.; Pierce, G.F.; Arruda, V.R.; Glader, B.; Ragni, M.; Rasko, J.J.; Kay, M.A. Successful transduction of liver in hemophilia by AAV-Factor IX and limitations imposed by the host immune response. Nat. Med. 2006, 12, 342–347. [Google Scholar] [CrossRef]
  129. Mingozzi, F.; Maus, M.V.; Hui, D.J.; Sabatino, D.E.; Murphy, S.L.; Rasko, J.E.J.; Ragni, M.V.; Manno, C.S.; Sommer, J.; Jiang, H.; et al. CD8+ T-cell responses to adeno-associated virus capsid in humans. Nat. Med. 2007, 13, 419–422. [Google Scholar] [CrossRef]
  130. Day, J.W.; Mendell, J.R.; Mercuri, E.; Finkel, R.S.; Strauss, K.A.; Kleyn, A.; Tauscher-Wisniewski, S.; Tukov, F.F.; Reyna, S.P.; Chand, D.H. Clinical Trial and Postmarketing Safety of Onasemnogene Abeparvovec Therapy. Drug Saf. 2021, 44, 1109–1119. [Google Scholar] [CrossRef]
  131. Tang, Y.; Fakhari, S.; Huntemann, E.D.; Feng, Z.; Wu, P.; Feng, W.Y.; Lei, J.; Yuan, F.; Excoffon, K.J.; Wang, K.; et al. Immunosuppression reduces rAAV2.5T neutralizing antibodies that limit efficacy following repeat dosing to ferret lungs. Mol. Ther. Methods Clin. Dev. 2023, 29, 70–80. [Google Scholar] [CrossRef]
  132. Ertl, H.C.J. Mitigating Serious Adverse Events in Gene Therapy with AAV Vectors: Vector Dose and Immunosuppression. Drugs 2023, 83, 287–298. [Google Scholar] [CrossRef]
  133. Meliani, A.; Boisgerault, F.; Hardet, R.; Marmier, S.; Collaud, F.; Ronzitti, G.; Leborgne, C.; Verdera, H.C.; Sola, M.S.; Charles, S.; et al. Antigen-selective modulation of AAV immunogenicity with tolerogenic rapamycin nanoparticles enables successful vector re-administration. Nat. Commun. 2018, 9, 4098. [Google Scholar] [CrossRef]
  134. Chandler, L.C.; Barnard, A.R.; Caddy, S.L.; Patrício, M.I.; McClements, M.E.; Fu, H.; Rada, C.; MacLaren, R.E.; Xue, K. Enhancement of Adeno-Associated Virus-Mediated Gene Therapy Using Hydroxychloroquine in Murine and Human Tissues. Mol. Ther. Methods Clin. Dev. 2019, 14, 77–89. [Google Scholar] [CrossRef]
  135. Li, X.; Wei, X.; Lin, J.; Ou, L. A versatile toolkit for overcoming AAV immunity. Front. Immunol. 2022, 13, 991832. [Google Scholar] [CrossRef]
  136. Elmore, Z.C.; Oh, D.K.; Simon, K.E.; Fanous, M.M.; Asokan, A. Rescuing AAV gene transfer from neutralizing antibodies with an IgG-degrading enzyme. JCI Insight 2020, 5, e139881. [Google Scholar] [CrossRef]
  137. Leborgne, C.; Barbon, E.; Alexander, J.M.; Hanby, H.; Delignat, S.; Cohen, D.M.; Mingozzi, F. IgG-cleaving endopeptidase enables in vivo gene therapy in the presence of anti-AAV neutralizing antibodies. Nat. Med. 2020, 26, 1096–1101. [Google Scholar] [CrossRef]
  138. Arjomandnejad, M.; Sylvia, K.; Blackwood, M.; Nixon, T.; Tang, Q.; Muhuri, M.; Gruntman, A.M.; Gao, G.; Flotte, T.R.; Keeler, A.M. Modulating immune responses to AAV by expanded polyclonal T-regs and capsid specific chimeric antigen receptor T-regulatory cells. Mol. Ther. Methods Clin. Dev. 2021, 23, 490–506. [Google Scholar] [CrossRef]
  139. Crudele, J.M.; Finn, J.D.; Siner, J.I.; Martin, N.B.; Niemeyer, G.P.; Zhou, S.; Mingozzi, F.; Lothrop, J.C.D.; 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. [Google Scholar] [CrossRef]
  140. Kumar, S.R.; Hoffman, B.E.; Terhorst, C.; de Jong, Y.P.; Herzog, R.W. The balance between CD8+ T Cell-mediated clearance of AAV-encoded antigen in the liver and tolerance is dependent on the vector dose. Mol. Ther. 2017, 25, 880–891. [Google Scholar] [CrossRef]
  141. Nathwani, A.C.; Reiss, U.M.; Tuddenham, E.G.; Rosales, C.; Chowdary, P.; McIntosh, J.; Della Peruta, M.; Lheriteau, E.; Patel, N.; Raj, D.; et al. Long-term safety and efficacy of factor IX gene therapy in hemophilia B. N. Engl. J. Med. 2014, 371, 1994–2004. [Google Scholar] [CrossRef]
  142. Subramanian, M.; McIninch, J.; Zlatev, I.; Schlegel, M.K.; Kaittanis, C.; Nguyen, T.; Agarwal, S.; Racie, T.; Alvarado, M.A.; Wassarman, K.; et al. RNAi-mediated rheostat for dynamic control of AAV-delivered transgenes. Nat. Commun. 2023, 14, 1970. [Google Scholar] [CrossRef]
  143. Pan, X.; Yue, Y.; Boftsi, M.; Wasala, L.P.; Tran, N.T.; Zhang, K.; Duan, D. Rational engineering of a functional CpG-free ITR for AAV gene therapy. Gene Ther. 2022, 29, 333–345. [Google Scholar] [CrossRef]
  144. Bertolini, T.B.; Shirley, J.L.; Zolotukhin, I.; Li, X.; Kaisho, T.; Xiao, W.; Kumar, S.R.P.; Herzog, R.W. Effect of CpG Depletion of Vector Genome on CD8+ T Cell Responses in AAV Gene Therapy. Front. Immunol. 2021, 12, 672449. [Google Scholar] [CrossRef]
  145. 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. Investig. 2013, 123, 2994–3001. [Google Scholar] [CrossRef]
  146. Haberman, R.P.; McCown, T.J.; Samulski, R.J. Novel Transcriptional regulatory signals in the adeno-associated virus terminal repeat junction element. J. Virol. 2000, 74, 8732–8739. [Google Scholar] [CrossRef]
  147. Colon-Moran, W.; Baer, A.; Lamture, G.; Stapleton, J.T.; Fischer, J.W.; Bhattarai, N. A short hepatitis C virus NS5A peptide expression by AAV vector modulates human T cell activation and reduces vector immunogenicity. Gene Ther. 2022, 29, 616–623. [Google Scholar] [CrossRef]
  148. Ran, G.; Chen, X.; Xie, Y.; Zheng, Q.; Xie, J.; Yu, C.; Pittman, N.; Qi, S.; Yu, F.-X.; Agbandje-McKenna, M.; et al. Site-Directed Mutagenesis Improves the Transduction Efficiency of Capsid Library-Derived Recombinant AAV Vectors. Mol. Ther. Methods Clin. Dev. 2020, 17, 545–555. [Google Scholar] [CrossRef]
  149. Frederick, A.; Sullivan, J.; Liu, L.; Adamowicz, M.; Lukason, M.; Raymer, J.; Luo, Z.; Jin, X.; Rao, K.N.; O’Riordan, C.R. Engineered Capsids for Efficient Gene Delivery to the Retina and Cornea. Hum. Gene Ther. 2020, 31, 756–774. [Google Scholar] [CrossRef]
  150. Bryant, D.H.; Bashir, A.; Sinai, S.; Jain, N.K.; Ogden, P.J.; Riley, P.F.; Kelsic, E.D. Deep diversification of an AAV capsid protein by machine learning. Nat. Biotechnol. 2021, 39, 691–696. [Google Scholar] [CrossRef]
  151. Mnyandu, N.; Limani, S.W.; Arbuthnot, P.; Maepa, M.B. Advances in designing Adeno-associated viral vectors for development of anti-HBV gene therapeutics. Virol. J. 2021, 18, 247. [Google Scholar] [CrossRef]
  152. Paulk, N.K.; Pekrun, K.; Zhu, E.; Nygaard, S.; Li, B.; Xu, J.; Chu, K.; Leborgne, C.; Dane, A.P.; Haft, A.; et al. Bioengineered AAV Capsids with Combined High Human Liver Transduction In Vivo and Unique Humoral Seroreactivity. Mol. Ther. 2018, 26, 289–303. [Google Scholar] [CrossRef]
  153. Barnes, C.; Scheideler, O.; Schaffer, D. Engineering the AAV capsid to evade immune responses. Curr. Opin. Biotechnol. 2019, 60, 99–103. [Google Scholar] [CrossRef]
  154. 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. [Google Scholar] [CrossRef]
  155. Bentler, M.; Hardet, R.; Ertelt, M.; Rudolf, D.; Kaniowska, D.; Schneider, A.; Vondran, F.W.; Schoeder, C.T.; Delphin, M.; Lucifora, J.; et al. Modifying immune responses to adeno-associated virus vectors by capsid engineering. Mol. Ther. Methods Clin. Dev. 2023, 30, 576–592. [Google Scholar] [CrossRef]
  156. Zhong, L.; Li, B.; Mah, C.S.; Govindasamy, L.; Agbandje-McKenna, M.; Cooper, M.; Herzog, R.W.; Zolotukhin, I.; Warrington, K.H., Jr.; Aken, K.A.W.-V.; et al. Next generation of adeno-associated virus 2 vectors: Point mutations in tyrosines lead to high-efficiency transduction at lower doses. Proc. Natl. Acad. Sci. USA 2008, 105, 7827–7832. [Google Scholar] [CrossRef] [PubMed]
  157. Martino, A.T.; Basner-Tschakarjan, E.; Markusic, D.M.; Finn, J.D.; Hinderer, C.; Zhou, S.; Ostrov, D.A.; Srivastava, A.; Ertl, H.C.J.; Terhorst, C.; et al. Engineered AAV vector minimizes in vivo targeting of transduced hepatocytes by capsid-specific CD8+ T cells. Blood 2013, 121, 2224–2233. [Google Scholar] [CrossRef] [PubMed]
  158. Hui, D.J.; Basner-Tschakarjan, E.; Chen, Y.; Davidson, R.J.; Buchlis, G.; Yazicioglu, M.; Mingozzi, F. Modulation of CD8+ T cell responses to AAV vectors with IgG-derived MHC class II epitopes. Mol. Ther. 2013, 21, 1727–1737. [Google Scholar] [CrossRef]
  159. Chuah, M.K.; Petrus, I.; De Bleser, P.; Le Guiner, C.; Gernoux, G.; Adjali, O.; Nair, N.; Willems, J.; Evens, H.; Rincon, M.Y.; et al. Liver-specific transcriptional modules identified by genome-wide in silico analysis enable efficient gene therapy in mice and non-human primates. Mol. Ther. 2014, 22, 1605–1613. [Google Scholar] [CrossRef]
  160. Fader, K.A.; Pardo, I.D.; Kovi, R.C.; Somps, C.J.; Wang, H.H.; Vaidya, V.S.; Ramaiah, S.K.; Sirivelu, M.P. Circulating neurofilament light chain as a promising biomarker of AAV-induced dorsal root ganglia toxicity in nonclinical toxicology species. Mol. Ther. Methods Clin. Dev. 2022, 25, 264–277. [Google Scholar] [CrossRef]
  161. Buss, N.; Lanigan, L.; Zeller, J.; Cissell, D.; Metea, M.; Adams, E.; Higgins, M.; Kim, K.H.; Budzynski, E.; Yang, L.; et al. Characterization of AAV-mediated dorsal root ganglionopathy. Mol. Ther. Methods Clin. Dev. 2022, 24, 342–354. [Google Scholar] [CrossRef]
  162. Hordeaux, J.; Buza, E.L.; Jeffrey, B.; Song, C.; Jahan, T.; Yuan, Y.; Wilson, J.M. MicroRNA-mediated inhibition of transgene expression reduces dorsal root ganglion toxicity by AAV vectors in primates. Sci. Transl. Med. 2020, 12, eaba9188. [Google Scholar] [CrossRef]
  163. Chan, Y.K.; Wang, S.K.; Chu, C.J.; Copland, D.A.; Letizia, A.J.; Costa Verdera, H.; Church, G.M. Engineering adeno-associated viral vectors to evade innate immune and inflammatory responses. Sci. Transl. Med. 2021, 13, eabd3438. [Google Scholar] [CrossRef]
  164. Reichel, F.F.; Dauletbekov, D.L.; Klein, R.; Peters, T.; Ochakovski, G.A.; Seitz, I.P.; Wilhelm, B.; Ueffing, M.; Biel, M.; Wissinger, B.; et al. AAV8 Can Induce Innate and Adaptive Immune Response in the Primate Eye. Mol. Ther. 2017, 25, 2648–2660. [Google Scholar] [CrossRef]
  165. Shao, W.; Earley, L.F.; Chai, Z.; Chen, X.; Sun, J.; He, T.; Deng, M.; Hirsch, M.L.; Ting, J.; Samulski, R.J.; et al. Double-stranded RNA innate immune response activation from long-term adeno-associated virus vector transduction. JCI Insight 2018, 3, e120474. [Google Scholar] [CrossRef] [PubMed]
  166. Wu, T.; Töpfer, K.; Lin, S.-W.; Li, H.; Bian, A.; Zhou, X.Y.; High, K.A.; Ertl, H.C. Self-complementary AAVs induce more potent transgene product-specific immune responses compared to a single-stranded genome. Mol. Ther. 2012, 20, 572–579. [Google Scholar] [CrossRef]
  167. Brown, D.; Altermatt, M.; Dobreva, T.; Chen, S.; Wang, A.; Thomson, M.; Gradinaru, V. Deep Parallel Characterization of AAV Tropism and AAV-Mediated Transcriptional Changes via Single-Cell RNA Sequencing. Front. Immunol. 2021, 12, 730825. [Google Scholar] [CrossRef]
  168. Hösel, M.; Lucifora, J.; Michler, T.; Holz, G.; Gruffaz, M.; Stahnke, S.; Büning, H. Hepatitis B virus infection enhances susceptibility toward adeno-associated viral vector transduction in vitro and in vivo. Hepatology 2014, 59, 2110–2120. [Google Scholar] [CrossRef]
  169. Bartolo, L.; Tong, S.L.C.; Chappert, P.; Urbain, D.; Collaud, F.; Colella, P.; Richard, I.; Ronzitti, G.; Demengeot, J.; Gross, D.A.; et al. Dual muscle-liver transduction imposes immune tolerance for muscle transgene engraftment despite preexisting immunity. JCI Insight 2019, 4, e127008. [Google Scholar] [CrossRef]
  170. Poupiot, J.; Verdera, H.C.; Hardet, R.; Colella, P.; Collaud, F.; Bartolo, L.; Davoust, J.; Sanatine, P.; Mingozzi, F.; Richard, I.; et al. Role of Regulatory T Cell and Effector T Cell Exhaustion in Liver-Mediated Transgene Tolerance in Muscle. Mol. Ther. Methods Clin. Dev. 2019, 15, 83–100. [Google Scholar] [CrossRef] [PubMed]
  171. Xu, D.; Fu, J.; Jin, L.; Zhang, H.; Zhou, C.; Zou, Z.; Zhao, J.-M.; Zhang, B.; Shi, M.; Ding, X.; et al. Circulating and liver resident CD4+CD25+ regulatory T cells actively influence the antiviral immune response and disease progression in patients with hepatitis B. J. Immunol. 2006, 177, 739–747. [Google Scholar] [CrossRef] [PubMed]
  172. Chen, J.; Wu, Q.; Yang, P.; Hsu, H.C.; Mountz, J.D. Determination of specific CD4 and CD8 T cell epitopes after AAV2-and AAV8-hF. IX gene therapy. Mol. Ther. 2006, 13, 260–269. [Google Scholar] [CrossRef]
  173. 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 in vivo hepatic gene transfer. J. Clin. Investig. 2003, 111, 1347–1356. [Google Scholar] [CrossRef]
  174. Franco, L.M.; Sun, B.; Yang, X.; Bird, A.; Zhang, H.; Schneider, A.; Brown, T.; Young, S.P.; Clay, T.M.; Amalfitano, A.; et al. Evasion of immune responses to introduced human acid α-glucosidase by liver-restricted expression in glycogen storage disease type II. Mol. Ther. 2005, 12, 876–884. [Google Scholar] [CrossRef] [PubMed]
  175. Chen, C.-C.; Sun, C.-P.; Ma, H.-I.; Fang, C.-C.; Wu, P.-Y.; Xiao, X.; Tao, M.-H. Comparative study of anti-hepatitis B virus RNA interference by double-stranded adeno-associated virus serotypes 7, 8, and 9. Mol. Ther. 2009, 17, 352–359. [Google Scholar] [CrossRef] [PubMed]
  176. Wörner, T.P.; Bennett, A.; Habka, S.; Snijder, J.; Friese, O.; Powers, T.; Agbandje-McKenna, M.; Heck, A.J.R. Adeno-associated virus capsid assembly is divergent and stochastic. Nat. Commun. 2021, 12, 1642. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Commonly used tools in HBV gene therapy development. To silence HBV gene expression, the RNAi pathway is activated exogenously by artificial microRNAs (miRNAs), for example. A microRNA duplex is incorporated into an RNA silencing complex (RISC) before the selection of one strand that will guide the RISC to the target HBV messenger RNAs. This results in HBV RNA degradation or translation suppression. Transcription activator-like endonucleases (TALENs) and zinc finger nucleases (ZFNs) work in pairs and require a right TALEN or right ZFN and left TALEN or left ZFN, conjugated to a cleavage domain, for double-strand cleavage to occur. TALENs consist of tandem repeats comprising 33–35 amino acids, whereas ZFNs consist of only 3–6 ZFs. The single guide RNA (sgRNA) associate and directs Cas 9 to the DNA sequence of interest. The Cas 9 enzyme creates double-stranded DNA breaks. Once cleavage has occurred, a double-strand DNA break occurs and recruits host machinery for repair. The error-prone non-homologous end joining (NHEJ) repair pathway is favored, which leads to insertions and deletions (indels) within the cccDNA/rcDNA (created with Biorender.com, accessed on 29 November 2023).
Figure 1. Commonly used tools in HBV gene therapy development. To silence HBV gene expression, the RNAi pathway is activated exogenously by artificial microRNAs (miRNAs), for example. A microRNA duplex is incorporated into an RNA silencing complex (RISC) before the selection of one strand that will guide the RISC to the target HBV messenger RNAs. This results in HBV RNA degradation or translation suppression. Transcription activator-like endonucleases (TALENs) and zinc finger nucleases (ZFNs) work in pairs and require a right TALEN or right ZFN and left TALEN or left ZFN, conjugated to a cleavage domain, for double-strand cleavage to occur. TALENs consist of tandem repeats comprising 33–35 amino acids, whereas ZFNs consist of only 3–6 ZFs. The single guide RNA (sgRNA) associate and directs Cas 9 to the DNA sequence of interest. The Cas 9 enzyme creates double-stranded DNA breaks. Once cleavage has occurred, a double-strand DNA break occurs and recruits host machinery for repair. The error-prone non-homologous end joining (NHEJ) repair pathway is favored, which leads to insertions and deletions (indels) within the cccDNA/rcDNA (created with Biorender.com, accessed on 29 November 2023).
Microorganisms 11 02985 g001
Figure 2. Development of AAVs for anti-HBV gene delivery. The wild-type AAV rep and cap ORFs are replaced with the anti-HBV sequence, most preferably driven by a liver-specific promoter (LSP). For primary microRNA (pri-miRNA) or short hairpin RNA (shRNA) and zinc finger nucleases (ZFNs), the self-complementary AAVs (scAAVs) are commonly used, whereas single-stranded AAVs (ssAAVs) are used for transcription activator-like endonucleases (TALENs) and RNA-guided clustered regulatory interspaced short palindromic repeats (CRISPR) and CRISPR associated (Cas) sequences. sgRNA expression is commonly driven by a polymerase III non-liver-specific promoter (NLSP). The AAV-mediated delivery of RNAi activators targeting Surface and X ORFs has been well characterized. AAVs carrying ZNFs targeting Surface, Polymerase, and Core ORFs have been reported. Although dual ssAAVs are promising for TALEN delivery, no anti-HBV TALEN expressing AAV has been reported. AAVs have been used to deliver CRISPR/Cas sequences against all HBV ORFs. The colored arrows on the HBV genome indicate the four open HBV reading frames, with all seven start codons and encoded proteins indicated (created with Biorender.com, accessed on 29 November 2023).
Figure 2. Development of AAVs for anti-HBV gene delivery. The wild-type AAV rep and cap ORFs are replaced with the anti-HBV sequence, most preferably driven by a liver-specific promoter (LSP). For primary microRNA (pri-miRNA) or short hairpin RNA (shRNA) and zinc finger nucleases (ZFNs), the self-complementary AAVs (scAAVs) are commonly used, whereas single-stranded AAVs (ssAAVs) are used for transcription activator-like endonucleases (TALENs) and RNA-guided clustered regulatory interspaced short palindromic repeats (CRISPR) and CRISPR associated (Cas) sequences. sgRNA expression is commonly driven by a polymerase III non-liver-specific promoter (NLSP). The AAV-mediated delivery of RNAi activators targeting Surface and X ORFs has been well characterized. AAVs carrying ZNFs targeting Surface, Polymerase, and Core ORFs have been reported. Although dual ssAAVs are promising for TALEN delivery, no anti-HBV TALEN expressing AAV has been reported. AAVs have been used to deliver CRISPR/Cas sequences against all HBV ORFs. The colored arrows on the HBV genome indicate the four open HBV reading frames, with all seven start codons and encoded proteins indicated (created with Biorender.com, accessed on 29 November 2023).
Microorganisms 11 02985 g002
Figure 3. Treg response modulation by AAV and HBV in the liver. Liver infection with AAV increases IL-10 production by KCs and enhances Foxp3+CD4+ Tregs differentiation and expansion. Secreted IL-10 can also mediate the conversion of antigen-specific CD4+ T cells into Foxp3+CD4+ Tregs and inhibit CD4+ T cell activity. Hepatitis, caused by HBV infection, stimulates MHC-II expression by hepatocytes. This enhances antigen presentation to naïve CD4+ T cells and results in their activation and differentiation to form Foxp3+CD4+ Tregs (created with BioRender.com accessed on 21 November 2023).
Figure 3. Treg response modulation by AAV and HBV in the liver. Liver infection with AAV increases IL-10 production by KCs and enhances Foxp3+CD4+ Tregs differentiation and expansion. Secreted IL-10 can also mediate the conversion of antigen-specific CD4+ T cells into Foxp3+CD4+ Tregs and inhibit CD4+ T cell activity. Hepatitis, caused by HBV infection, stimulates MHC-II expression by hepatocytes. This enhances antigen presentation to naïve CD4+ T cells and results in their activation and differentiation to form Foxp3+CD4+ Tregs (created with BioRender.com accessed on 21 November 2023).
Microorganisms 11 02985 g003
Table 1. Approaches to circumventing AAV-mediated immunotoxicity.
Table 1. Approaches to circumventing AAV-mediated immunotoxicity.
TargetGoalStrategyReferences
Host
Deplete effecter T cells and induce Treg expansionCo-administration of AAV with rapamycin[133]
Inhibit TLR9 signalingHydroxychloroquine administration before AAV injection [134]
Inhibit complement activationEculizumab or APL-9 or IgG-degrading enzyme administration before AAV injection[29,118,119,135]
Induce toleranceInjecting AAV-specific CAR Treg following AAV injection [138]
Genome
Reduce doseEfficient transgene engineering[139]
Reduce dose and induce toleranceTransgene expression from weaker hepatocyte-specific promoters and RNAi-mediated transgene self-regulation[140,141,142]
Avoid TLR 9 signaling activationCpG island reduction or TLR9 inhibition via oligonucleotide incorporation in the genome [143,144,145]
Avoid dsRNA sensingEngineering of the expression cassette, such that transcription from 3′ITR is inhibited[146]
Repress the activation of primary and memory T-cell responsesCo-expressing NS5A with the transgene[147]
Capsid
Reduce dose, de-target APCs, and induce toleranceRational designs and directed evolution to generate capsids with high hepatocyte transduction efficiencies [148,149,150,151]
Avoid complement activationRational designs and directed evolution to generate capsids that will evade pre-existing antibodies[135,152,153,154]
Avoid TLR9 signaling activationTLR9 inhibiting peptides insertion in the capsid[155]
Reduce antigen presentationMutation of surface-exposed tyrosine residues [156,157]
Enhance Treg proliferation and reduce CD8+ T cell activityIncorporation of the MHC II epitopes derived from IgG in the AAV capsid[158]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Jacobs, R.; Dogbey, M.D.; Mnyandu, N.; Neves, K.; Barth, S.; Arbuthnot, P.; Maepa, M.B. AAV Immunotoxicity: Implications in Anti-HBV Gene Therapy. Microorganisms 2023, 11, 2985. https://doi.org/10.3390/microorganisms11122985

AMA Style

Jacobs R, Dogbey MD, Mnyandu N, Neves K, Barth S, Arbuthnot P, Maepa MB. AAV Immunotoxicity: Implications in Anti-HBV Gene Therapy. Microorganisms. 2023; 11(12):2985. https://doi.org/10.3390/microorganisms11122985

Chicago/Turabian Style

Jacobs, Ridhwaanah, Makafui Dennis Dogbey, Njabulo Mnyandu, Keila Neves, Stefan Barth, Patrick Arbuthnot, and Mohube Betty Maepa. 2023. "AAV Immunotoxicity: Implications in Anti-HBV Gene Therapy" Microorganisms 11, no. 12: 2985. https://doi.org/10.3390/microorganisms11122985

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop