Skip to main content
Springer logoLink to Springer
. 2017 May 31;40(4):497–517. doi: 10.1007/s10545-017-0053-3

Gene therapy for monogenic liver diseases: clinical successes, current challenges and future prospects

Julien Baruteau 1,2,3,, Simon N Waddington 3,4, Ian E Alexander 5,6, Paul Gissen 1,2,7
PMCID: PMC5500673  PMID: 28567541

Abstract

Over the last decade, pioneering liver-directed gene therapy trials for haemophilia B have achieved sustained clinical improvement after a single systemic injection of adeno-associated virus (AAV) derived vectors encoding the human factor IX cDNA. These trials demonstrate the potential of AAV technology to provide long-lasting clinical benefit in the treatment of monogenic liver disorders. Indeed, with more than ten ongoing or planned clinical trials for haemophilia A and B and dozens of trials planned for other inherited genetic/metabolic liver diseases, clinical translation is expanding rapidly. Gene therapy is likely to become an option for routine care of a subset of severe inherited genetic/metabolic liver diseases in the relatively near term. In this review, we aim to summarise the milestones in the development of gene therapy, present the different vector tools and their clinical applications for liver-directed gene therapy. AAV-derived vectors are emerging as the leading candidates for clinical translation of gene delivery to the liver. Therefore, we focus on clinical applications of AAV vectors in providing the most recent update on clinical outcomes of completed and ongoing gene therapy trials and comment on the current challenges that the field is facing for large-scale clinical translation. There is clearly an urgent need for more efficient therapies in many severe monogenic liver disorders, which will require careful risk-benefit analysis for each indication, especially in paediatrics.

Introduction

The liver is a key-regulator of multiple complex metabolic pathways and the hepatocyte is a primary cell type affected in numerous inherited genetic/metabolic diseases (Clayton 2002). Despite a wide range of disease-specific conventional therapies, liver replacement therapies remain a valid strategy and even a potential cure for many monogenic liver disorders due to the ability to restore the defective pathway (Sokal 2006). Liver replacement options include whole or partial organ (Spada et al 2009), or hepatocytes transplantation (Dhawan et al 2006). The shortage of donors, the associated mortality/morbidity and need for immunosuppression, however, often limit this option to severely affected patients and those aged more than 3 months or weighing greater than 5 kg (Haberle et al 2012). In the past decade, liver-directed gene therapy has emerged as a promising alternative to transplantation in monogenic liver disorders.

Overview of gene therapy development: reaching maturity

Gene therapy, by providing additional functional gene copies, has been considered for decades as an attractive option for treatment of monogenic disorders (Wirth et al 2013). According to the Gartner hype cycle, a graphical representation depicting the maturity of novel technologies, gene therapy reached its “peak of inflated expectation” in the mid-1990s which was paralleled by a rapid rise in clinical trial activity and the publication of early proof-of-concept studies for genetic and acquired conditions such as adenosine-deaminase deficiency (ADA-SCID) (Blaese et al 1995; Bordignon et al 1995) and brain tumours, respectively (Puumalainen et al 1998). This period of inflated expectation was critiqued in the Orkin-Motulsky report commissioned by the National Institute of Health (Orkin and Motulsky 1995). While acknowledging the extraordinary promise of gene therapy, the report emphasised the need for greater focus on gene transfer technology and the basic science of gene transfer. Soon after, the field plunged into its “trough of disillusionment” following the death of a young adult, Jesse Gelsinger, in a clinical trial for ornithine transcarbamylase (OTC) deficiency (Raper et al 2003). Optimism arising from the subsequent clinical success in the treatment of X-linked severe combined immunodeficiency (SCID-X1) (Cavazzana-Calvo et al 2000) was soon dampened by the occurrence of leukaemia in five out of 20 patients secondary to insertional mutagenesis (Hacein-Bey-Abina et al, 2003a, b; Fischer et al 2010; Mukherjee and Thrasher 2013), causing the death of one participant (Mukherjee and Thrasher 2013). Resultant concerns over gene therapy were further compounded by growing awareness of the challenges imposed by vector-induced immune responses (Mingozzi and High 2007). Disbelief and doubt followed, leading to a decline in financial investment (Ledley et al 2014). In parallel, these adverse events motivated researchers to seek a better understanding of the challenges posed by disease pathophysiology and to develop safer and more efficient vectors. Recent clinical successes in various inherited orphan diseases such as Leber’s congenital amaurosis (Bainbridge et al 2008; Cideciyan et al 2008; Maguire et al 2008), X-linked adrenoleukodystrophy (Cartier et al 2009), metachromatic leukodystrophy (Biffi et al 2013) and haemophilia B (Nathwani et al 2011) and the first market authorisation granted by the European Medicines Agency (EMA) in 2012, to Glybera® for lipoprotein lipase deficiency (Bryant et al 2013), are driving the field up the “slope of enlightenment” and onto the “plateau of productivity”. As a result of this success, various biotechnology companies dedicated to gene therapy development have been created and received substantial financial investment (Cassiday 2014). In parallel, the number of gene therapy-based clinical trials has risen rapidly in recent years (http://www.abedia.com/wiley/(accessed 2017 Jan 06); Ginn et al 2013).

Strategies for hepatocyte-directed gene transfer

A growing toolbox is available for gene transfer, which has been the preferred approach in recent human trials targeting hepatocytes. Various elements in the choice of transgene expression cassette design, mode of delivery and the subset of patients targeted influence the efficacy of gene therapy (Fig. 1).

Fig. 1.

Fig. 1

The triad to consider for successful gene therapy

Parameters of vector delivery

The mode of transgene delivery is crucial: i) local injection allows highly selective expression, but in a limited area. Conversely an intravenous injection allows a broad distribution balanced by non-specificity. Injections in the hepatic artery or the portal vein improve the selectivity but require cannulation with its associated risks (Fumoto et al 2013). Peripheral intravenous delivery provides similar transduction compared to intrahepatic or intraportal routes for AAV vectors (Sarkar et al 2006; Nathwani et al 2007); ii) higher doses of vector achieve greater transduction, but may generate more severe immune responses (Raper et al 2003; Mingozzi and High 2013).

Host pre-sensitisation or acquired immune responses

The immune response against the vector and/or the transgene product might preclude the expected therapeutic effect (Jooss and Chirmule 2003; Zaiss and Muruve 2008; Wold and Toth 2013).

Immune memory of pre-exposure to wild-type viruses can prevent efficient hepatocyte transduction by pre-existing neutralising antibodies and might account for differences in the severity of immune responses observed after systemic injection. Pre-immunisation against the transgene product can occur when the recombinant transgenic protein has been administered. For example, in haemophilia B patients, this can result in the generation of anti-factor IX antibodies when treated by recombinant factor IX (Armstrong et al 2014). Accordingly prior immunisation needs to be carefully considered in clinical trial enrolment criteria.

Acquired immune responses after systemic gene delivery are common. Innate immune responses are triggered by antigen presenting cells such as dendritic cells or macrophages initiating the release of proinflammatory cytokines (interleukins 1 and 6, tumour necrosis factor α (TNFα), type I interferon α and β) via stimulation of Toll-like receptors (TLR). Specific and long-lasting antigen-specific immune responses are mediated by B- and T-cells and involve secretion of neutralising antibodies and CD8+ cytotoxic T lymphocytes, respectively, regulated by the recruitment of helper and regulatory CD4+ T cells. Whatever the viral vector considered, immune responses share various similarities, and must always be carefully considered (Bessis et al 2004; Tang et al 2006; Annoni et al 2013; Calcedo and Wilson 2013; Basner-Tschakarjan and Mingozzi 2014).

Design and selection of the gene transfer vector

The transgene expression cassette contains i) a transgene, which is commonly a cDNA and may be codon-optimised to achieve higher expression of the transgene product, ii) an enhancer/promoter, the selection of which determines the level of transgene expression, cell-type restricted specificity of expression and also influences the risk of insertional mutagenesis, iii) various pre- and/or post-regulatory elements to stabilise transgene mRNA and therefore increase the yield of transgene product, e.g. addition of an intron downstream of the promoter containing a bacterial replication origin (Lu et al 2017) or the woodchuck hepatitis virus post-transcriptional regulatory element (WPRE) (Lipshutz et al 2003), respectively.

Several options are available to deliver the transgene expression cassette to the target cell/organ

Injection of naked DNA, either as plasmids (Doenecke et al 2010; Oishi et al 2016) or mini-circles (Viecelli et al 2014; Hou et al 2016; Wu et al 2016), is a simple mode of transgene delivery which lends itself to local delivery and is relatively non-immunogenic compared to some viral vector approaches (Wolff and Budker 2005). Mini-circles are devoid of plasmid backbone DNA; this may enhance transgene expression by overcoming heterochromatin formation and avoiding inflammation triggered by bacterial DNA (Mayrhofer et al 2009). These approaches allow easy production of therapeutic material with a good safety profile and capable of eliciting long-lasting transgene expression in post-mitotic tissues (Wolff and Budker 2005; Kay et al 2010). These approaches have been employed in ∼17% of gene therapy trials so far (http://www.abedia.com/wiley/(accessed 2017 Jan 06)). The hydrodynamic injection technique, developed in small animal models, consists of injecting DNA plasmids or mini-circles in a large vehicle volume to flood the liver with pressurised DNA solution; this disrupts vascular endothelium, and allows high levels of transgene expression in small or large animal models (Liu et al 1999). Although hydrodynamic injections are difficult to translate to humans, intravascular hydrodynamic procedures with partial catheterisation for liver-directed gene delivery have shown some success in large animals (Sendra et al 2016; Yokoo et al 2016). However, current translatable options of non-viral approaches remain limited. Therefore viral vectors, acting as Trojan horses to increase transduction efficiency, are frequently considered.

Non-viral vectors are synthetically produced biological particles, in which the transgene is encapsulated, or complexed, and released at the target site. Various engineered nanoparticles exist, e.g. liposomes or/and polymers (Chira et al 2015). These options have several advantages: easy production, no restriction of the transgene size, and a reliable safety profile. Limitations are the stability of these particles, cellular uptake and a limited ability to achieve long-lasting transgene expression (Elsabahy et al 2011). These are therefore suboptimal delivery vehicles for liver-directed clinical trials.

Virus-derived vectors represent an attractive approach based on their relatively efficient transduction human cells. The main vectors that have been used in clinical trials are derived from adenoviruses (21%), retroviruses (excluding lentiviruses) (19%), adeno-associated viruses (7%) and lentiviruses (6%) (http://www.abedia.com/wiley/(accessed 2017 Jan 06)).

Retroviral vectors

Gamma-retroviruses such as murine leukaemia viruses (MLVs) are RNA viruses encoding gag, pol and env genes flanked by long terminal repeats (LTRs), which carry enhancers/promoter elements and are required for integration. After transduction of the target cell, reverse-transcription generates a double-stranded DNA copy of the proviral genome, which then integrates in the host genome providing long-term transgene expression. Gamma-retroviruses are unable to transduce non-dividing cells as the nuclear membrane prevents retroviral vectors from entering the nucleus (Miller et al 1990). This explains why this vector is more often considered for ex vivo gene therapy in which cultured target cells are stimulated to replicate and then transduce. Lentiviruses are a class of retroviruses, the most widely known of which is HIV1. Lentiviral vectors are able to transduce dividing and non-dividing cells, which broadens their application. Retroviral vectors have relatively large transgene capacities (7.5 kilobases (kb)) (Verma and Somia 1997). In fact payloads exceeding 14 kilobases have been packaged into lentiviral vectors (Counsell et al, 2017).

Clinical applications and limitations: In an early gene therapy trial, a γ-retroviral vector was used in an ex vivo approach with autologous hepatocytes in five patients with homozygous familial hypercholesterolemia. This showed a mild improvement of lipid profiles in two patients with a very low rate of stable engraftment at 4 months after gene therapy (Grossman et al 1994; Grossman et al 1995). Several limitations emerged from this trial: i) the need for two invasive procedures, i.e. the surgical resection of a liver lobe to obtain sufficient primary hepatocytes for transduction as hepatocytes cannot be expanded in culture and reinjection of transduced hepatocytes into the portal circulation via a local catheter, with the associated risks of venous thrombosis, catheter misplacement and haemorrhage (Grossman et al 1994; Raper et al 1996). The efficiency with which hepatocytes were harvested and transduced was low, 30% and 10% respectively (Grossman et al 1995). Protocols for lentiviral-mediated gene therapy have improved with in vitro transduction efficacy reaching 90% in non-human and human hepatocytes (Nguyen et al 2009). Hepatocyte transplantation, however, remains relatively inefficient and variable, likely due to poor engraftment, limited persistence of engrafted hepatocytes and the lack of a proliferative advantage (Gramignoli et al 2015).

Retroviral-mediated in vivo gene therapy was well tolerated when vector was administered intravenously in haemophilia A patients, but no significant clinical benefits were observed (Powell et al 2003). Preclinical studies failed to accurately predict the therapeutic dose and it was suggested that retroviral vectors were unable to transduce non-dividing hepatocytes (Chuah et al 2004). Therefore, alternative viral vectors, including lentiviral vectors, have been developed for treatment of hemophilias. Indeed two promising approaches with lentiviral vectors are in preclinical development: i) systemic lentiviral-mediated liver-restricted gene therapy in a dog model of haemophilia B, which showed long-term efficacy, induction of liver tolerogenic properties in stimulating CD4+CD25+FoxP3+ regulatory T cells and no evidence of genotoxicity in mice (Cantore et al 2015). The immune tolerance is of particular interest for patients with anti-FIX inhibitors, who are currently excluded from gene therapy trials; ii) ex vivo transduction of haematopoietic stem cells (HSC) in mice with successful FIX gene expression to target cells of the erythroid (Chang et al 2008) or the megakaryocyte lineage (Chen et al 2014). A lentiviral vector is currently in preclinical development for haemophilia B (Dolgin 2016). Lentiviral vectors are able to accommodate large transgenes such as FVIII gene for haemophilia A (Kuether et al 2012). Lentiviral-mediated liver-directed ex vivo gene therapy has been successfully reported in a pig model of tyrosinaemia 1 using the selective advantage of Fah +/+ modified hepatocytes (Hickey et al 2016).

The risk of insertional mutagenesis has been reported with retroviral vectors (Cavazzana-Calvo et al 2000; Mukherjee and Thrasher 2013), however, there is evidence that the risk is lower with lentiviral vectors (Kotterman et al 2015). To improve safety, self-inactivating (SIN) vectors have been developed in which LTR enhancer/promoter elements in the U3 region have been deleted (Miyoshi et al 1998; Zufferey et al 1998). This, however, does not completely eliminate the risk of insertional mutagenesis as heterologous enhancer-promoter elements still need to be included in vector constructs. Genotoxicity has been observed after foetal injections of non-primate and primate SIN-lentiviral vectors (Nowrouzi et al 2013; Condiotti et al 2014). So far, more than 125 patients over 14 years have been treated with haematopoietic stem cells or T cells transduced by lentiviral vectors with no oncogenic event reported (Cartier et al 2009; Biffi et al 2013; McGarrity et al 2013; Booth et al 2016). Another approach relies on the mutation of the integrase protein to generate non-integrating or integration-deficient lentiviral vectors (Nightingale et al 2006; Philippe et al 2006; Yanez-Munoz et al 2006), which have shown long-lasting gene expression in non-dividing tissues (Apolonia et al 2007; Rahim et al 2009) and phenotype correction in mice with haemophilia B (Suwanmanee et al 2014).

Adenoviral vectors

Adenoviruses are non-enveloped double-stranded DNA viruses with a large 36 kb genome and are capable of transducing dividing and non-dividing cells. In humans, a common target is epithelial cells of the respiratory or gastrointestinal tracts causing mild upper respiratory tract infection, gastroenteritis, or asymptomatic seroconversion. More than 55 serotypes are described but most vectors are derived from endemic serotypes 2 and 5 (Piccolo and Brunetti-Pierri 2014). The seroprevalence against the most common serotype (Adenovirus 5) is high with neutralising antibodies in 45–80% (Kotterman et al 2015). Adenoviruses elicit a sustained innate and cytotoxic-mediated immunity, which leads to the clearance of transduced cells (Tang et al 2006; Thaci et al 2011). Neither adenoviruses nor adenoviral vectors have been associated with genotoxicity in humans (Stephen et al 2010). Adenoviral vectors can accommodate large transgenes, which remain episomal in the transduced cell but enable a long-lasting transgene expression in quiescent or dividing cells. In animal models, adenoviral vectors exhibit strong liver tropism through interaction with coagulation factor X (Kalyuzhniy et al 2008; Vigant et al 2008; Waddington et al 2008).

Clinical applications and limitations

In adenoviruses, expression of genes occurs in two early and late phases. The early phase is mediated by E1-E4 transcription units. Proteins encoded by E1 are essential for viral gene expression and DNA replication. Late gene expression is mediated by an internal promoter (Benihoud et al 1999). First-generation adenoviral vectors have E1 or E1-E3 regions removed and are theoretically replication-defective. However, these vectors keep a mild “leaky” expression of viral genes and are still able to synthesise some viral proteins, likely due to E1-like protein in the target cells (Zhang et al 1998; Lozier et al 1999). Moreover, the immunogenic properties of first-generation adenoviral vectors cause severe acute innate and chronic adaptive immune responses in small and large animal models (Yang et al 1994). To reduce this immune response, deletion of transcription units E1/E2/E3, E1/E4/E3, E1/E2/E3/E4) has been introduced (Alba et al 2005). Only the last combination achieved a reduction of vector toxicity, but with a reduced duration of the transgene expression (Gao et al 1996; Raper et al 1998; Andrews et al 2001). A second generation (E1- and E4- deleted) adenoviral 5 vector was used for the OTC deficiency trial, in which Jesse Gelsinger, a young adult with late-onset OTC deficiency enrolled in the highest dose group, died after developing a fatal acute toxic reaction with fulminant inflammatory response and multi-organ failure hours after the injection of the vector (Raper et al 2003). It has been hypothesised that this was caused by an innate immune response with a cytokine storm triggered by antigen presenting cells against capsid proteins (Raper et al 2003). The reason for severity of this immune response remains unclear as another patient injected with the same dose exhibited only mild flu-like symptoms. A genetic predisposition or an immune memory response caused by pre-exposure to adenoviruses might partly explain this discrepancy (Wilson 2009). In this trial, safety issues were surprisingly not dose-related and had not been predicted to this extent by animal studies (Raper et al 2002). Furthermore, no significant clinical benefit was observed (Raper et al 2002).

To limit this immune response, “gutless” or helper-dependent adenoviral vectors (HD-Ad) have been designed by deletion of all coding regions except the ITRs and the packaging signal (Ψ) required for the encapsidation of the adenoviral genome, which have been replaced by the transgene cassette and stuffer DNA (Alba et al 2005). This new generation of vectors have shown an improved safety profile, exhibiting a reduced acute innate immune response and an absence of chronic toxicity. In first-generation adenoviral vectors, the “leaky” expression of the remaining viral genes have a direct cytotoxic effect and triggers an adaptive cellular immune response directed against the transduced cell, which in turn results in transient transgene expression and chronic toxicity (Brunetti-Pierri and Ng 2011). As HD-Ad do not contain any viral genes, this late toxicity is not observed, which allows a long-term transgene expression as observed in small (Kim et al 2001; Toietta et al 2005) and large animal models (Morral et al 1999; Brunetti-Pierri et al 2006). This has allowed successful long-term phenotypic correction of various liver monogenic disorders (Brunetti-Pierri and Ng 2011), among which haemophilia A (Reddy et al 2002; Hu et al 2011), haemophilia B (Ehrhardt and Kay 2002), OTC deficiency (Mian and Lee 2002), glycogen storage disease 1A (Koeberl et al 2007), Criggler-Najjar syndrome (Schmitt et al 2014), primary hyperoxaluria type 1 (Castello et al 2016), acute intermittent porphyria (Unzu et al 2013), phenylketonuria (Cerreto et al 2012), familial hypercholesterolaemia caused by mutations in the LDLR (Nomura et al 2004) and ApoE genes (Belalcazar et al 2003) in rodents, haemophilia A (Brown et al 2004; McCormack et al 2006) and B (Ehrhardt et al 2003; Brunetti-Pierri et al 2005), glycogen storage disease type 1 (Crane et al 2012) in dogs, and haemophilia A and B and Pompe disease (Rastall et al 2016) in non human primates.

However, the acute innate immune response directed against capsid proteins is not abolished in HD-Ad vectors (Muruve et al 2004). For both first-generation and helper-dependent adenoviral vectors, this acute immune response is dose-dependent and can be lethal at high doses in non human primates (Morral et al 2002; Brunetti-Pierri et al 2004). Differences in the severity of the immune response between species have emerged due to variable interactions between blood cells and hepatic microarchitecture such as size of liver sinusoidal fenestration (Piccolo and Brunetti-Pierri 2014). The activation of this innate immunity is multifactorial. Adenoviral particles can trigger the immune response by binding to Toll-like receptors (TLR2, TLR9) at the surface of antigen presenting cells, and/or activate the complement cascade in the bloodstream (Kiang et al 2006; Zhu et al 2007). Kupffer cells recognise the adenoviral capsid either via antibody-mediated opsonisation or in binding complement factors. Kupffer cells develop a pro-inflammatory state with necrotic death, which further disseminate the immune response (Schiedner et al 2003).

In vivo HD-Ad mediated gene therapy has been performed in one patient in a phase I trial for haemophilia A. After a single intravenous low-dose injection, the patient developed flu-like symptoms with transient fever, chills, back pain, headache and transient biological abnormalities including thrombocytopenia, laboratory features of disseminated intravascular coagulopathy, increase in interleukin 6 levels, and elevated liver transaminase levels peaking at 7 days (marked as grade 3 liver toxicity) (Chuah et al 2004; White and Monahan 2005; Chandler and Venditti 2016). The patient expressed 1% FVIII for some months but the trial was halted for safety reason although biological abnormalities came back to normal within 19 days. Unfortunately, this trial has not been published in a peer-reviewed format and few details are available (Piccolo and Brunetti-Pierri 2014). The cause of these symptoms remains unclear although, as supposed for the OTC trial involving Jesse Gelsinger, the innate immune response against the adenoviral capsid and its subsequent release of cytokines has been suspected (Chuah et al 2004). Contamination by an adenoviral helper virus remains a possible explanation (Chuah et al 2004).

Despite limited application for liver monogenic disorders, adenoviral vectors have been successfully used for oncolytic virotherapy (Rosewell Shaw and Suzuki 2016) and vaccination (Majhen et al 2014), which exploits adenoviral immunogenicity.

Adeno-associated viral vectors

Adeno-associated viruses (AAV) are non-enveloped, single-stranded DNA viruses that belong to the Dependovirus genus and the Parvoviridae family. Initially identified as a contaminant of an adenoviral preparation (Atchison et al 1965), the virus was later shown to require co-infection with a helper virus to replicate. In the absence of helper virus, AAV can enter target cells and establish latent infection through genomic integration and/or formation of episomes. AAV is widely considered non-pathogenic and has yet to be definitively linked to disease causation. AAVs can transduce dividing and non-dividing cells. The seroprevalence against the most common serotype AAV2 is 40–60% (Louis Jeune et al 2013). AAV virions consist of an icosahedral capsid of approximately 22 nm in diameter enclosing a 4.7 kb single-stranded genome. The genome is flanked by two 145 nucleotide inverted terminal repeats (ITRs) containing all of the necessary cis-acting functions for proviral rescue, genome replication and packaging. The viral genome encodes 4 Rep proteins required for proviral rescue and genome replication, and three viral proteins VP1, VP2 and VP3, which assemble to form the capsid (Fig. 2A) (Samulski and Muzyczka 2014). Some, but not all, AAV capsid serotypes (Earley et al 2017) require expression of an assembly-activating protein (AAP) encoded by an alternative reading frame of the Cap gene and providing scaffolding activity (Naumer et al 2012). Numerous AAV serotypes have been isolated from humans, non-human primates and other species, with the viral capsid determining species and target cell tropism through interaction with a diversity of cell surface receptors/co-receptors and intracellular trafficking pathways that remain incompletely understood. A multi-serotype AAV receptor has been recently identified (Pillay et al 2016), but its precise role in uptake and trafficking has yet to be elucidated (Summerford and Samulski 2016). For example, AAV3B uses the human hepatocyte growth factor (hHGF) receptor, which restricts transduction to primates and especially to the liver (Vercauteren et al 2016).

Fig. 2.

Fig. 2

Synthesis of an AAV vector. (a) Initially, the single-stranded proviral DNA is excised to remove Rep and Cap genes from different wild type AAV serotypes. The transgene expression cassette containing the promoter, the transgene and various regulatory elements is cloned between the 2 ITRs, which are the only wild type AAV sequences retained. (b) For vector synthesis, triple transfection of three plasmids is performed in a packaging cell with proviral plasmid encoding the recombinant viral genome, a plasmid containing Rep and Cap and a helper plasmid. “Pseudotyped” AAV vectors contain ITRs from a specific AAV serotype (usually AAV2) and a Cap gene encoding viral proteins (VP1, 2 and 3) from a different serotype (e.g AAV8) in order to provide organ-specific transduction of the recombinant AAV vector named AAV2/8. AAV: adeno-associated virus; Adv: adenovirus; ITR: inverted terminal repeat

Since 2004, AAV vectors have emerged as the leading candidates for gene therapy in monogenic liver disorders with the best accepted benefit-risk ratio (Dolgin 2016). Therefore, the following sections focus on this gene transfer approach detailing clinical successes and current limitations.

AAV2 is the most widely studied serotype and was the first to be vectorized. To generate a recombinant AAV vector the native Rep and Cap genes are removed and replaced by a transgene expression cassette with only the flanking ITRs retained (Fig. 2A). Recombinant virus is produced by supplying Rep and Cap and necessary adenoviral helper functions in trans. A major development in AAV vector technology was the demonstration that recombinant AAV2 genomes can be cross-packaged, or pseudo-serotyped, with the capsids from other AAV serotypes (Rabinowitz et al 2002). This has dramatically broadened the cell types that can be efficiently targeted with AAV vectors. For example, pseudo-serotyping a recombinant AAV2 vector genome with the AAV8 capsid (designated AAV2/8) enhances tropism for hepatocytes, particularly in the mouse (Fig. 2B). AAV vectors bind to target cells via specific receptors and co-receptors that differ in a capsid-dependent manner and are taken up by endocytosis or macropinocytosis, before being trafficked to the nucleus for capsid uncoating. The uncoated genomes can remain in the nucleus in single-stranded form, be converted to double-stranded episomes or undergo genomic integration (Fig. 3) (Berry and Asokan 2016). Conversion of input single-stranded genomes to double-stranded transcriptionally active forms occurs with variable efficiency in different cell types. Self-complementary (sc) vectors differ from single-stranded vectors (ss) in that they contain a self-complementary transgene cassette that folds back on itself to form double-stranded DNA thereby bypassing the requirement for second strand synthesis, which is considered as a rate-limiting step for transgene expression. As a consequence the packaging size of the transgene cassette in scAAV is reduced by half (McCarty 2008).

Fig. 3.

Fig. 3

AAV vector uptake, in-cell processing and initiation of the immune response. Fenestrated endothelium of hepatic sinusoids allows the AAV vector to freely reach the hepatocyte. Once reaching the target cell, the vector binds an extracellular receptor and co-receptor specific to the capsid motifs. After an uptake by endocytosis, the vector is trafficked in the cytoplasm in early then late endosome. Acidification of the endosome modifies the capsid conformation. After endosomal escape, the AAV vector enters the nucleus via the nuclear pore complex. Capsid uncoating and release of the proviral DNA precede the synthesis of the 2nd strand of DNA. The viral genome then persists either as a non-integrated single- or double-stranded episome (99%) or (small percentage) integrates into the host genome (1%). Expression of the transgene is followed by synthesis of the protein of interest. Cell-mediated immune responses are initiated by the degradation of capsid or the transgene product (protein) in the proteasome and presentation at the surface of the transduced cell via the major histocompatibility complex I. CD8+ T cells recognise the antigen at the cell surface and initiate the immune cascade. Neutralising antibodies bind to the vector in the bloodstream and impair or prevent successful transduction of the organ target. MHC1: major histocompatibility complex I

Clinical successes of liver-directed AAV-mediated gene therapy

A rapidly increasing number of publications have reported proof-of-concept for AAV-based gene therapy in animal models for various inherited liver disorders including urea cycle defects, organic acidurias, phenylketonuria, glycogen storage disease type Ia, long chain fatty acid oxidation disorders, homozygous familial hypercholesterolemia, primary hyperoxaluria type I and progressive familial intrahepatic cholestasis (Hastie and Samulski 2015; Junge et al 2015).

In parallel, pioneering trials have been conducted since the 2000s, two of which targeted haemophilia B. This disease is an attractive target for gene therapy as an increase in plasma factor IX (FIX) of as little as 1% can confer significant phenotypic improvement. Haemophilia B is a burden in public healthcare systems with an annual cost of $300,000 for severely affected patients (Angelis et al 2015).

In 2004, a ssAAV2.ApoE/hAAT.hFIX vector, administered via the hepatic artery, showed a transient increase of plasma FIX from <1% to 3–11% over 4 weeks followed by a gradual decline over 4–8 weeks concomitant with transient asymptomatic rise in transaminases levels (Manno et al 2006), later recognised as T cell-mediated cytotoxicity (Mingozzi et al 2007).

In 2009, Nathwani et al, injected a scAAV2/8.LP1.hcoFIX vector via a peripheral intravenous route and elicited a long-lasting (>5 years) increase of plasma FIX from <1% to 1–8% (Nienhuis et al 2016). Elevated transaminases occurring 7–10 weeks post-injection resolved after an oral course of corticosteroids, but were associated with a decrease of 50–70% in plasma FIX levels attributable to a cellular immune response against capsid epitopes (Nathwani et al 2014).

D’Avola et al, recently reported results of a trial of scAAV2/5.hAAT.hcoPBGD vector in acute intermittent porphyria with peripheral intravenous delivery. No vector-related safety issues were reported and the rate of disease-related hospitalisation decreased, potentially as a consequence of closer metabolic follow-up. No change was observed in the levels of metabolic biomarkers (D’Avola et al 2016). This might be explained by a less efficient liver transduction mediated by AAV5 relative to AAV8 and a reduced expression of the episomal transgene compared to the endogenous gene of interest (Baruteau et al 2017). However, no liver biopsy data was available to address this assumption.

Preliminary reports from ongoing clinical trials have confirmed Nathwani’s promising results for haemophilia B. After a single intravenous injection of AAV vectors with different capsids encoding the FIX gene or its Padua FIX variant, which contains a gain-of-function mutation, reported stabilised plasma FIX levels have ranged from 3 to 8% in the AMT-060 trial sponsored by Uniqure (Miesbach et al 2016) and the DTX-101 sponsored by Dimension Therapeutics (http://dimensiontx.com) to 20–44% in the high-dose cohort of the BAX 335 trial sponsored by Shire (Monahan et al 2015) and in the SPK-9001 trial sponsored by Spark Therapeutics/Pfizer (George et al 2016). In a haemophilia A gene therapy trial, BioMarin reported plasma factor VIII (FVIII) from 4 to 60% in the high-dose group of the BMN 270 trial (Pasi et al 2016) (Table 1). Importantly, endogenous FVIII is primarily secreted by endothelial cells (Fahs et al 2014). All the AAV-based trials have so far involved only adult patients, who had an undetectable baseline titre of neutralising antibodies to the capsid (usually accepted cut-off of 1/5 serum dilution). Monogenic liver disorders in AAV-based gene therapy development pipelines of pharmaceutical companies include OTCD, glycogen storage disease type Ia, citrullinemia type I, phenylketonuria, Wilson disease, methylmalonic acideamia and Crigler-Najjar syndrome (Kattenhorn et al 2016).

Table 1.

Clinical trials of gene therapy products for liver monogenic disorders. Intraportal and intrahepatic routes of administration relate to injection on the portal vein and the hepatic artery respectively. FVIII: factor VIII; FIX: factor IX; HD-adenovirus: helper-dependent adenovirus; LDL: low density lipoprotein; MoML: Moloney murine leukaemia virus; NA: not applicable; OTC: ornithine transcarbamylase; PBGD: porphobilinogen deaminase. Information sources: clinicaltrial.gov (accessed 06/01/2017) and company websites. If the date of the start of the trial was not available in clinicaltrials.gov website, the date of publication of the results is mentioned. The list of trials announced for 2017 is indicative and does not pretend to be exhaustive

Year (start of trial) Viral vector Disease Product Therapeutic Sponsor Route Dose Number of treated patients Status Reference Clinicaltrial.gov identifier
1992 MoMLV Retrovirus 5 Homozygous familial hypercholesterolaemia NA LDL receptor gene University of Michigan Ann Arbour Intraportal (Ex vivo approach) 1 to 3.3 × 10e9 hepatocytes 5 Terminated Grossman et al 1994
Grossman et al 1995
Raper et al 1996
NCT00004809
1998 Adenovirus 5 Ornithine transcarbamylase deficiency NA OTC gene University of Pennsylvania Intrahepatic 2 × 10e9 to 6 × 10e11vg/kg 18 Terminated Raper et al 2003 NCT00004498
1999 AAV2 Haemophilia B NA Factor IX gene NA Intramuscular 2 × 10e11 to 1.8 × 10e12vg/kg 8 Terminated Kay et al 2000
Manno et al 2003
NA
2003 MoMLV Retrovirus Haemophilia A NA FVIII gene Chiron Intravenous 2.8 × 10e7 to 4.4 × 10e8vg/kg 13 Terminated Powell et al 2003 NA
2004 HD-Adenovirus Haemophilia A NA FVIII gene GenStar Intravenous 4.3 × 10e10vg/kg 1 Terminated Chuah et al 2004 NA
2004 AAV2 α1-antitrypsin NA hAAt University Massachussets Intramuscular 2.1 × 10e12 to 6.9 × 10e13vg 12 Terminated Flotte et al 2004
Brantly et al 2006
NCT00377416
2004 AAV2 Haemophilia B NA Factor IX gene Avigen Intrahepatic 8 × 10e10 to 2 × 10e12vg/kg 7 Terminated Manno et al 2006 NCT00076557
2006 AAV1 α1-antitrypsin NA hAAT University Massachussets Intramuscular 6.9 × 10e12 to 6 × 10e13vg 9 Terminated Brantly et al 2009 NCT00430768
2009 AAV2/8 Haemophilia B NA Factor IX gene St Jude Children’s Hospital Intravenous 2 × 10e11 to 2 × 10e12vg/kg 10 Recruiting Nathwani et al 2014 NCT00979238
2010 AAV1 α1-antitrypsin NA hAAT Applied Genetic Technologies Intramuscular 6 × 10e12 to 6 × 10e12vg/kg 9 Terminated Flotte et al 2011 NCT01054339
2012 AAV2/8 Haemophilia B BAX 335 Padua mutant factor IX gene Shire Intravenous 2 × 10e11 to 3 × 10e12vg/kg 6 Terminated Monahan et al 2015 NCT01687608
2014 AAV2/5 Acute intermittent porphyria NA PBGD gene Digna Biotech Intravenous 5 × 10e11 to 1.8 × 10e13vg/kg 8 Terminated D’Avola et al 2016 NCT02082860
2015 Engineered AAV Haemophilia B SPK-9001 Padua mutant factor IX gene Spark Therapeutics Intravenous 5 × 10e11 vg/kg (Low dose) 7 Recruiting George et al 2016 NCT02484092
AAV2/5 Haemophilia B AMT-060 Factor IX gene Uniqure Biopharma Intravenous 5 × 10e12 to 2 × 10e13vg/kg 5 Recruiting Miesbach et al 2016 NCT02396342
AAV2/rh10 Haemophilia B DTX101 Factor IX gene Dimension Therapeutics Intravenous 1.6 × 10e12 to 1 × 10e13vg/kg Undisclosed Recruiting clinicaltrials.gov NCT02618915
AAV2/5 Haemophilia A BMN 270 Factor VIII gene BioMarin Pharmaceuticals Intravenous 6 × 10e12 to 6 × 10e13vg/kg 9 Recruiting Pasi et al 2016 NCT02576795
2016 AAV2/6 Haemophilia B SB-FIX Factor IX integrating in the albumin locus via Zinc-finger-nuclease Sangamo Bioscines Intravenous cDNA 4 × 10e12 to 4 × 10e13vg/kg and ZFN 5 × 10e11 to 5 × 10e12vg/kg Undisclosed Recruiting clinicaltrials.gov NCT02695160
AAV2/8 Homozygous familial hypercholesterolaemia RGX-501 LDL receptor gene University of Pennsylvania/Regenxbio Intravenous 2.5 × 10e12 to 7.5 × 10e12vg/kg Undisclosed Recruiting clinicaltrials.gov NCT02651675
Announced for 2017 AAV2/8 Ornithine transcarbamylase deficiency DTX301 OTC gene Dimension Therapeutics Intravenous 2 × 10e12 to 2 × 10e13vg/kg Undisclosed Recruiting clinicaltrials.gov NCT02991144
Engineered AAV Haemophilia A SPK-8011 Factor VIII gene Spark Therapeutics Intravenous Undisclosed Undisclosed Recruiting clinicaltrials.gov NCT03003533
AAV2/8 Haemophilia A GO-8 Factor VIII gene University College London Intravenous 6 × 10e11 to 6 × 10e12vg/kg NA Not yet recruiting clinicaltrials.gov NCT03001830
Engineered AAV Haemophilia B FLT-180 Undisclosed Freeline Therapeutics
AAV2/8 Haemophilia A BAX-888 Factor VIII gene Shire
AVV Haemophilia A DTX201 Factor VIII gene Dimension Therapeutics/Bayer
AAV2/6 Haemophilia A SB-525 Factor VIII integrating in the albumin locus via Zinc-finger-nuclease Sangamo Biosciences
AAV2/8 Mucopolysaccharidosis VI MeuSIX ARSB gene NA
AAV2/8 Crigler Najjar AT342 UGT1A1 gene Audentes Therapeutics

Current challenges

Insertional mutagenesis

Despite more than 170 AAV-based human trials approved, ongoing or completed (http://www.abedia.com/wiley/(accessed 2017 Jan 06)), no tumorigenic events have been reported so far. AAV vector genome mainly persists as episome in the transduced cell with a relatively low proportion of vector genomes undergoing integration preferentially in transcriptionally active genes, damaged DNA or enriched CpG islands (McCarty et al 2004). Experiments in neonatal mice have identified an increased risk of hepatocellular carcinoma (HCC) after systemic injection. This risk increased with the enhancer/promoter activity, younger age at time of injection and vector dose (Donsante et al 2007; Chandler et al 2015; Chandler et al 2017). Analysis of integration sites identified a rodent-specific hotspot in the Rian locus. Integration studies from human trials have not shown such hotspots, but rather a genome wide integration pattern involving neither HCC-related genes nor the human Rian homologue, Dlk1-Dio3 (Kaeppel et al 2013; Gil-Farina and Schmidt 2016).

Controversies remain regarding the possible insertional mutagenic effects of wild type AAV. Detection of a clonal expansion of wild type AAV2 sequences in 11/193 HCCs within HCC-related genes (Nault et al 2015) initiated a passionate and unresolved debate about “driver” or “passenger” cancer-related genetic modifications (Berns et al 2015; Buning and Schmidt 2015). The cumulative safety experience with the rapidly growing number of AAV-based trials targeting the human liver, combined with the low rate of HCC-associated AAV integrations despite the high seroprevalence of wild type AAV in the human population (e.g. >50% for AAV2) (Thwaite et al 2015) are consistent with a favourable safety profile of AAV vectors. Nevertheless the findings of Nault et al warrant further studies and mandate close monitoring in ongoing human trials.

Immune response

After vector delivery, non-specific innate immunity triggers both type I interferon signalling involved in transgene silencing (Suzuki et al 2013) and the release of proinflammatory cytokines (Jayandharan et al 2011). Highly-specific and long-lasting adaptive immunity generates B- and T-cell responses against the capsid and/or the transgene product (Fig. 3). Neutralising antibodies against the capsid, even at low titers, inhibit transduction after systemic delivery (Jiang et al, 2006a, b). This barrier is of substantial concern to gene therapy development and the ongoing liver-directed trials are recruiting only seronegative patients without neutralising antibodies against the AAV capsid. This narrows the target population as the seroprevalence against liver-specific AAV serotypes ranges from 20 to 30% for AAV5, 6 and 8 to 50–60% for AAV2 (Louis Jeune et al 2013). Cross-reactivity between serotypes is commonly >50% (Boutin et al 2010). This seroprevalence varies depending on geographic origin (Calcedo et al 2009) and age. Neonates receive maternal antibodies by transplacental transfer and acquired with maternal milk, which are lost over the first months of life. Thereafter, seroprevalence remains negligible until 3 years of age after which the seroconversion rate progressively increases until adulthood (Calcedo et al 2011; Li et al 2012).

Human CD8+ T-cell mediated immune responses are involved in AAV hepatotoxicity and were initially encountered during the first haemophilia B trial (Manno et al 2006). Capsid epitopes, presented via the major histocompatibility complex I (MHC1), were shown to drive expansion of a pre-existing pool of CD8+ memory T cells acquired during a previous co-infection of wild type AAV and helper virus (adenovirus or herpes virus for example). This response was dose-dependent (Mingozzi and High 2013) and could be stimulated by alternate capsids (Mingozzi et al 2007).

Hepatic tolerogenic properties involve the proliferation of a specific T cell subset, CD4+CD25+FoxP3+ Treg cells (Cooper et al 2009) interacting with Kupffer cells (Breous et al 2009). Expansion of these cells suppresses cytotoxic immunity against AAV-transduced hepatocytes and induces immunotolerance, e.g. after neonatal injection (Shi et al 2013). Regulatory tolerance requires continuous antigen presentation and has been successfully induced with transgenic proteins (Shi et al 2013; Perrin et al 2016). In contrast, capsid proteins are rapidly eliminated in the proteasome (Berry and Asokan 2016) and therefore very unlikely to induce tolerance.

Various approaches aim to overcome these unwanted immune responses in order either to treat seropositive patients or to prevent sensitisation against the AAV capsid, which would allow reinjection in the future. Capsid modification targeting specific epitopes can evade host immunity (Tseng and Agbandje-McKenna 2014). Any strategy optimising the transduction of the target organ such as optimised expression cassette design, or capsid modifications will in turn decrease the amount of vector required for a similar effect. The decrease in vector dose will further reduce the immune response (Mingozzi and High 2011). Various protocols involving transient immunosuppression have been proposed in large animal models and humans. These include plasmapheresis (Monteilhet et al 2011; Chicoine et al 2014), monoclonal antiCD20 antibody (rituximab) (Mingozzi et al 2013; Corti et al 2014), non-depleting antiCD4 antibody (McIntosh et al 2012), sirolimus (Corti et al 2014), cyclosporine A (McIntosh et al 2012; Mingozzi et al 2012), tacrolimus with mycophenolate mofetil (Chicoine et al 2014), proteasome inhibitors, e.g. bortezomib (Monahan et al 2010) and corticosteroids (Flanigan et al 2013; Chicoine et al 2014; Nathwani et al 2014). One currently controversial strategy to blunt anti-capsid immune responses, is to co-inject “full” AAV vectors and “empty” capsids as decoys, in an attempt to competitively bind existing antibodies (Tse et al 2015).

Optimised targeting

The ideal AAV vector would exclusively transduce the desired target cells. This would limit unwanted immune responses, avoid ectopic transgene expression and further reduce the already low risk of germline transmission. Extensive capsid-focused research based on high-throughput in vitro and in vivo screening of “libraries” of new capsid variants is ongoing in order to optimise vectors for specific applications. These “accelerated evolution” libraries are generated using strategies such as error-prone PCR (Kotterman and Schaffer 2014; Deverman et al 2016) and “capsid shuffling” with random cut-paste sequences of wild type cap genes (Kay 2011; Louis Jeune et al 2013; Choudhury et al 2016). This approach is paying off by generating re-engineered AAV variants with increased transduction efficiency in primary human hepatocytes (Lisowski et al 2014).

Germline transmission

The risk for this phenomenon is difficult to quantify. Insertion of AAV sequences into the genome of a gamete could potentially interfere with normal foetal development or promote tumorigenicity in the progeny. So far adult patients enrolled in AAV trials with systemic delivery have been required to use contraception. Vector sequences have been detected transiently in semen of treated patients in AAV2 (Manno et al 2006) or AAV8 (Nathwani et al 2014) trials with the latest clearance of the vector observed at 12 weeks post-injection, quicker in younger men, but not in an AAV5 trial (D’Avola et al 2016). Vector was observed in seminal fluid but not in motile sperm and spermatogonia aligned with previous studies (Arruda et al 2001; Couto et al 2004).

Suboptimal animal models

The following examples demonstrate why animal experiments provide limited value for predicting effects in human trials:

  • T-cell mediated cytotoxicity observed in the first haemophilia B trial was not predicted by experiments in mice, dogs or non-human primates (Pien et al 2009). Unlike research animals, humans are exposed to wild type AAV infections generating anti-AAV memory T cells, which are reactivated at the time of vector exposure (Mingozzi et al 2007).

  • An over 80% rate of HCC was observed in a mouse model of methylmalonic acidaemia injected neonatally with AAV2/8. Integration occurred in the Rian hotspot, a rodent-specific locus absent in other vertebrate genomes (Chandler et al 2015).

  • The patterns of liver transgene expression in the hepatic lobule varies among different species. For example, using the AAV8 capsid, transgene expression is predominantly pericentral in mice and dogs and periportal in non-human primates (Fig. 4) (Bell et al 2011). This is of particular importance for liver diseases where metabolic zonation underpins that certain metabolic functions occur predominantly in certain areas of the liver lobule, which is the functional unit of the liver. For example, the urea cycle activity mostly takes place in periportal hepatocytes (Gebhardt and Matz-Soja 2014). To achieve adequate control of severe hyperammonaemia in the OTCD mouse model, therefore, requires a much higher than expected dose of AAV2/8 vector carrying OTC transgene, which might be explained by the non-physiological pattern of liver transduction (Cunningham et al 2011). Thus, it is difficult to reliably extrapolate vector doses for human translation from studies in mice in liver diseases with metabolic zonation like OTCD.

  • AAV2/8 vectors are capable of transducing 100% hepatocytes in adult mice (Cunningham et al 2008), but data from human trials in haemophilia B have shown an increase in plasma FIX only 2 to 8% (Nienhuis et al 2016), suggesting much less efficient AAV2/8-mediated hepatocyte transduction in humans. Interestingly, most studies in a chimeric FRG (Fah −/− /Rag2 −/− /Il2rg −/−) mouse-human liver model (Lisowski et al 2014; Wang et al 2015; Vercauteren et al 2016) showed that AAV3B and AAV3B-derived vectors (AAV3-ST, AAV-LK03) are able to transduce human hepatocytes approximately 10 times more efficiently than AAV2/8 whilst transduction of murine hepatocytes is minimal (Lisowski et al 2014).

Fig. 4.

Fig. 4

Species-related differences in transduction of the hepatic lobule by AAV vector compared with metabolic zonation for ammonia clearance: example with AAV2/8 vector

The FRG mouse has a combination of tyrosinaemia type I and immunodeficiency phenotypes and is an attractive model to study human hepatocytes in vivo with the intention of overcoming limitations due to species-specificity (Azuma et al 2007). Human Fah +/+ hepatocytes have a selective growth advantage relative to the Fah-deficient native mouse hepatocytes allowing human cell engraftment up to 90% of the liver mass (Azuma et al 2007). Moreover, this model can address disease-specific questions if engrafted with hepatocytes from patients with liver specific disorders. Recently, the even more complex FRGN (Fah −/− /Rag2 −/− /Il2rg −/− /NOD) mouse model has been described in which FRG mice, developed in a non-obese diabetic (NOD) mouse strain, are simultaneously co-transplanted with human hepatocytes and human haematopoietic stem cells (Wilson et al 2014).

Limited capacity of AAV vectors

The single-stranded AAV vector can accommodate a transgene cassette of approximately 4.6 to 5 kb (Hirsch et al 2016). This capacity is reduced by half (2.3 kb) in self-complementary (i.e double-stranded) vectors. This is a major limitation compared to non-viral or other common viral vectors like lentiviral/retroviral vectors (up to at least 14 kb (Counsell et al 2017)) or helper-dependent adenoviruses (up to at least 38.9 kb (Suzuki et al 2011)). To deliver oversized transgenes, several approaches have been developed. Designing mini-promoters or mini-genes of interest can be successful (Yan et al 2015). Alternatively, dual AAV co-transduction has been successfully tested either with split AAV or fragment AAV (Hirsch et al 2016). Split AAVs use the inherent tendency for intermolecular genome association observed with AAV genomes via either homologous recombination (HR) or non-homologous end joining (NHEJ) to produce concatemers. The overlapping approach uses vectors A and B, which display a homology sequence to promote intermolecular homologous recombination (Duan et al 2001; Koo et al 2014). In the trans-splicing approach, two splice sites in 3′ cDNA of vector A and 5′ cDNA of vector B are recognised in concatemerized provirus to generate the single DNA molecule of the oversized gene of interest (Duan et al 2001). A combination of the two approaches is known as the hybrid trans-splicing technique (Trapani et al 2014). In fragment AAV, the transgene is not entirely encapsidated but only fragments of different size, which can recombine on overlapping regions (Hirsch et al 2016). A limitation common to these approaches is reduced functional transduction efficiency.

Limited manufacturing capability

For the last couple of years, the rise in demand for good-manufacturing practice (GMP) AAV vectors for preclinical and clinical studies has created a bottleneck, delaying a number of projects. The industry is progressively taking up the gauntlet and developing improved and innovative methods for vector production in larger bioreactors with optimised reagents, purification techniques and packaging cell lines (Clement and Grieger 2016; Grieger et al 2016).

Gene therapy requires an innovative economic model for success in modern healthcare

Patients with inherited metabolic disorders individually impose a much heavier financial burden on the healthcare system compared to the average person. For example, the lifetime cost for methylmalonic/propionic acidaemias and Gaucher disease is $1.5 and 5 million, respectively (Li et al 2015; Orkin and Reilly 2016). Gene therapy has the potential to achieve substantial savings. For example, a haemophilia B trial has shown that single injection of the gene therapy product in a cohort of ten patients can save more than $2.5 million over three years for the healthcare system in the UK (Nathwani et al 2014).

It is likely that to recover the investment in product development, companies will need pricing gene therapy treatments ambitiously when their products reach market. However, the cost of treatment will need to be affordable for public healthcare systems. The first gene therapy product approved by the European Medicines Agency (EMA), Glybera®, is marketed by Uniqure. This drug has been in development for 8 years, with the initial developer going bankrupt and a currently proposed market cost of $1.2 million per patient (Bryant et al 2013). GlaxoSmithKline‘s Strimvelis® is the second gene therapy product to reach the market and was approved by the EMA in 2016. This gene therapy product, which targets ADA-SCID has been in development for over 16 years and the proposed cost is $665,000 per patient (Hoggatt 2016; Schimmer and Breazzano 2016).

Over the last 20 years, greater than $4.3 billion have been spent on development of gene therapy technology and return on this investment is still awaited (Ledley et al 2014). Although most of the gene therapy programmes remain in the early stages of development, healthcare economists are generating models to cost treatments, which might provide lifelong cures. A pay-for-performance system has been proposed with yearly-capped annuity paid to the pharmaceutical company if criteria of a metabolic control of the disease are met (Touchot and Flume 2015). These criteria might reflect cost-effectiveness and not only cost-saving. This approach values the gain in quality of life estimated by quality-adjusted life years (QALY) analysis, which includes many parameters such as lifespan and ability to work (Orkin and Reilly 2016). Vouchers systems with longer financial incentives might be another option (Schimmer and Breazzano 2016).

In parallel, the regulatory framework is evolving with the progress of technology and the increasing experience being gathered from human trials. The Food and Drug Administration (FDA) and the EMA have published recommendations for gene therapy products (Narayanan et al 2014). The need for shorter and less expensive paths to clinical trials and conditional approval relying more on human data for safety and efficacy has now been recognised, as exemplified by the FDA’s Breakthrough Therapies programme and the EMA’s Adaptive Pathways and Priority Medicines (PRIME) schemes which were launched in 2012, 2014 and 2016, respectively (Mullard 2016). These more flexible pathways will need to be agreed by government funding bodies (Macaulay 2015).

Considerations for paediatric application

Paediatric administration of gene therapy has several theoretical advantages (McKay et al 2011). These include prevention of early death or irreversible neurological sequellae, transduction of stem/progenitor cells and possible avoidance of immune response as neonates have an immature immune system, while infants and children have lower rates of pre-existing anti-AAV immunity. The potential for even earlier gene therapy intervention has been explored in late gestation foetal macaques by intrahepatic injection of scAAV2/8 and scAAV2/5 vectors. Plasma human FIX levels of 8–112% were observed during a median follow-up period of 14 months without evidence hotspot integration and HCC (Mattar et al 2011).

There is a theoretical risk, however, of increased tumorigenicity in the developing liver as observed in experiments performed in neonatal mice (Donsante et al 2007; Chandler et al 2015). A further challenge in the paediatric liver is the likely progressive loss of vector genomes over time in concert with hepatocellular proliferation. More than 90% of the AAV-delivered transgene cassettes exists as non-integrated episomes (Cunningham et al 2008). The human liver weight doubles at 4 months, 16 months, 6 years and 12 years of age, which means that the adult liver is 16 times heavier than the neonatal liver (Coppoletta and Wolbach 1933; Sunderman and Boerner 1949). Therefore, it is unlikely that a neonatal injection will be sufficient to provide lifelong correction of the phenotype in metabolic liver diseases, with reinjection during the phase of rapid liver growth likely to be necessary.

An alternative approach to reinjection could be the use of integrating vectors and or locus-specific genome engineering. Sangamo Therapeutics Inc. is developing tools for transgene integration into the albumin locus and uses zinc finger nucleases coupled with AAV technology. Other genome editing tools have been successfully tested with AAV vectors in a neonatal OTC deficiency mouse model. In these experiments, two AAV vectors were injected simultaneously, one of which encoded the transgene and the other the enzymatic system for integration or site specific cutting by Piggybac transposase (Cunningham et al 2015) and CRISPR-Cas9 (Yang et al 2016), respectively.

Muscle-directed gene therapy for liver monogenic disorders

Muscle-directed gene therapy has been developed for liver monogenic disorders with secreted protein such as haemophilia A and B and α1-antitrypsin deficiency. Intramuscular injections might circumvent some caveats observed with systemic injection, the common route of delivery for liver-targeted gene therapy: limited biodistribution with reduced risk of germline transmission, minimal exposure to circulating neutralising antibodies, reduced dose of vector for a similar effect.

A proof of concept using AAV2 vectors in mice (Herzog et al 1997) and dogs (Herzog et al 1999) affected by haemophilia B paved the way for a clinical trial (Kay et al 2000; Manno et al 2003), which showed a safe profile with long-standing expression in some patients (Jiang et al, 2006a, b; Buchlis et al 2012) but only a mild increase in plasma factor IX around 1% (Manno et al 2003). Depending on the dose considered, dozens to hundreds of intramuscular injections are necessary, which makes this route particularly impractical. For instance, Manno et al administered between 10 to 90 injections per patients in lower limbs (Manno et al 2003). Similarly, a proof of concept in mice with haemophilia A has been reported (Mah et al 2003).

Three AAV-mediated clinical trials have been conducted for α1-antitrypsin deficiency. In this disorder, a plasma level of wild-type (M) α1-antitrypsin above 11 μM is considered reducing the risk of developing emphysema. A first trial based on an AAV2 capsid showed an acceptable safety profile with mild local reactions at the site of intramuscular injection (redness, tenderness, bruising) and a seroconversion against AAV2. Unfortunately, only one out of 12 patients demonstrated a minimal increase of plasma M α1-antitrypsin at 82 nM (Flotte et al 2004; Brantly et al 2006). Two other trials (phase I then phase II) were conducted with a vector based on AAV1 capsid known for its better muscle transduction compared to AAV2 (Flotte et al 2011). In both trials, minor side effects and a seroconversion against AAV1 were observed (Brantly et al 2009; Flotte et al 2011). A moderate infiltration of reactive T lymphocytes in muscle biopsies was noticed (Flotte et al 2011). Plasma levels of M α1-antitrypsin were still mild, although improving in the latest trial due to higher doses injected (412 to 694 nM in the highest dose group) but far from the targeted protective level of 11 μM (Flotte et al 2011).

Conclusion

Over the last decade, major discoveries in the understanding of viral vector biology have generated promising results in pioneering clinical trials for haemophilia B using AAV vectors. This has paved the way for a wider development of AAV-mediated gene therapy for monogenic liver disorders. Although several clinical, manufacturing and economic challenges remain, this approach to treatment for severely debilitating diseases generated widespread enthusiasm shared by clinicians, researchers and investors alike. Gene transfer technologies are reaching an exciting threshold of efficacy and promise to revolutionise the management of many currently untreatable diseases.

Compliance with ethical standards

Conflict of interest

J. Baruteau, S. N. Waddington, I. E. Alexander, and P. Gissen declare that they have no conflict of interest.

Funding

J.B. is funded by Action Medical Research for Children Charity (grant GN2137) and Great Ormond Street Hospital (GOSH) Children’s Charity Clinical Research Starter Grant. J.B., S.N.W. and P.G. are in receipt of the UK Medical Research Council (MRC) grant MR/N019075/1. S.N.W. is funded by the MRC grants MR/N026101/1 and MR/P026494/1.

Footnotes

Presented at the Annual Symposium of the Society for the Study of Inborn Errors of Metabolism, Rome, Italy, September 6–9, 2016

References

  1. Alba R, Bosch A, Chillon M. Gutless adenovirus: last-generation adenovirus for gene therapy. Gene Ther. 2005;12(Suppl 1):S18–S27. doi: 10.1038/sj.gt.3302612. [DOI] [PubMed] [Google Scholar]
  2. Andrews JL, Kadan MJ, Gorziglia MI, Kaleko M, Connelly S. Generation and characterization of E1/E2a/E3/E4-deficient adenoviral vectors encoding human factor VIII. Mol Ther. 2001;3:329–336. doi: 10.1006/mthe.2001.0264. [DOI] [PubMed] [Google Scholar]
  3. Angelis A, Tordrup D, Kanavos P. Socio-economic burden of rare diseases: a systematic review of cost of illness evidence. Health Policy. 2015;119:964–979. doi: 10.1016/j.healthpol.2014.12.016. [DOI] [PubMed] [Google Scholar]
  4. Annoni A, Goudy K, Akbarpour M, Naldini L, Roncarolo MG. Immune responses in liver-directed lentiviral gene therapy. Transl Res. 2013;161:230–240. doi: 10.1016/j.trsl.2012.12.018. [DOI] [PubMed] [Google Scholar]
  5. Apolonia L, Waddington SN, Fernandes C, et al. Stable gene transfer to muscle using non-integrating lentiviral vectors. Mol Ther. 2007;15:1947–1954. doi: 10.1038/sj.mt.6300281. [DOI] [PubMed] [Google Scholar]
  6. Armstrong EP, Malone DC, Krishnan S, Wessler MJ. Costs and utilization of hemophilia a and B patients with and without inhibitors. J Med Econ. 2014;17:798–802. doi: 10.3111/13696998.2014.953679. [DOI] [PubMed] [Google Scholar]
  7. Arruda VR, Fields PA, Milner R, et al. Lack of germline transmission of vector sequences following systemic administration of recombinant AAV-2 vector in males. Mol Ther. 2001;4:586–592. doi: 10.1006/mthe.2001.0491. [DOI] [PubMed] [Google Scholar]
  8. Atchison RW, Casto BC, Hammon WM. Adenovirus-associated defective virus particles. Science. 1965;149:754–756. doi: 10.1126/science.149.3685.754. [DOI] [PubMed] [Google Scholar]
  9. Azuma H, Paulk N, Ranade A, et al. Robust expansion of human hepatocytes in fah−/−/Rag2−/−/Il2rg−/− mice. Nat Biotechnol. 2007;25:903–910. doi: 10.1038/nbt1326. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Bainbridge JW, Smith AJ, Barker SS, et al. Effect of gene therapy on visual function in Leber’s congenital amaurosis. N Engl J Med. 2008;358:2231–2239. doi: 10.1056/NEJMoa0802268. [DOI] [PubMed] [Google Scholar]
  11. Baruteau J, Waddington SN, Alexander IE, Gissen P (2017) Delivering efficient liver-directed AAV-mediated gene therapy. Gene Ther. doi:10.1038/gt.2016.90 [DOI] [PubMed]
  12. Basner-Tschakarjan E, Mingozzi F. Cell-mediated immunity to AAV vectors, evolving concepts and potential solutions. Front Immunol. 2014;5:350. doi: 10.3389/fimmu.2014.00350. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Belalcazar LM, Merched A, Carr B, et al. Long-term stable expression of human apolipoprotein A-I mediated by helper-dependent adenovirus gene transfer inhibits atherosclerosis progression and remodels atherosclerotic plaques in a mouse model of familial hypercholesterolemia. Circulation. 2003;107:2726–2732. doi: 10.1161/01.CIR.0000066913.69844.B2. [DOI] [PubMed] [Google Scholar]
  14. Bell P, Wang L, Gao G, et al. Inverse zonation of hepatocyte transduction with AAV vectors between mice and non-human primates. Mol Genet Metab. 2011;104:395–403. doi: 10.1016/j.ymgme.2011.06.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Benihoud K, Yeh P, Perricaudet M. Adenovirus vectors for gene delivery. Curr Opin Biotechnol. 1999;10:440–447. doi: 10.1016/s0958-1669(99)00007-5. [DOI] [PubMed] [Google Scholar]
  16. Berns KI, Byrne BJ, Flotte TR, et al. Adeno-associated virus type 2 and hepatocellular carcinoma? Hum Gene Ther. 2015;26:779–781. doi: 10.1089/hum.2015.29014.kib. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Berry GE, Asokan A. Cellular transduction mechanisms of adeno-associated viral vectors. Curr Opin Virol. 2016;21:54–60. doi: 10.1016/j.coviro.2016.08.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Bessis N, GarciaCozar FJ, Boissier MC. Immune responses to gene therapy vectors: influence on vector function and effector mechanisms. Gene Ther. 2004;11(Suppl 1):S10–S17. doi: 10.1038/sj.gt.3302364. [DOI] [PubMed] [Google Scholar]
  19. Biffi A, Montini E, Lorioli L, et al. Lentiviral hematopoietic stem cell gene therapy benefits metachromatic leukodystrophy. Science. 2013;341:1233158. doi: 10.1126/science.1233158. [DOI] [PubMed] [Google Scholar]
  20. Blaese RM, Culver KW, Miller AD, et al. T lymphocyte-directed gene therapy for ADA-SCID: initial trial results after 4 years. Science. 1995;270:475–480. doi: 10.1126/science.270.5235.475. [DOI] [PubMed] [Google Scholar]
  21. Booth C, Gaspar HB, Thrasher AJ. Treating immunodeficiency through HSC Gene therapy. Trends Mol Med. 2016;22:317–327. doi: 10.1016/j.molmed.2016.02.002. [DOI] [PubMed] [Google Scholar]
  22. Bordignon C, Notarangelo LD, Nobili N, et al. Gene therapy in peripheral blood lymphocytes and bone marrow for ADA- immunodeficient patients. Science. 1995;270:470–475. doi: 10.1126/science.270.5235.470. [DOI] [PubMed] [Google Scholar]
  23. Boutin S, Monteilhet V, Veron P, et al. Prevalence of serum IgG and neutralizing factors against adeno-associated virus (AAV) types 1, 2, 5, 6, 8, and 9 in the healthy population: implications for gene therapy using AAV vectors. Hum Gene Ther. 2010;21:704–712. doi: 10.1089/hum.2009.182. [DOI] [PubMed] [Google Scholar]
  24. Brantly ML, Spencer LT, Humphries M, et al. Phase I trial of intramuscular injection of a recombinant adeno-associated virus serotype 2 alphal-antitrypsin (AAT) vector in AAT-deficient adults. Hum Gene Ther. 2006;17:1177–1186. doi: 10.1089/hum.2006.17.1177. [DOI] [PubMed] [Google Scholar]
  25. Brantly ML, Chulay JD, Wang L et al (2009) Sustained transgene expression despite T lymphocyte responses in a clinical trial of rAAV1-AAT gene therapy. Proc Natl Acad Sci USA 106:16363–16368 [DOI] [PMC free article] [PubMed]
  26. Breous E, Somanathan S, Vandenberghe LH, Wilson JM. Hepatic regulatory T cells and Kupffer cells are crucial mediators of systemic T cell tolerance to antigens targeting murine liver. Hepatology. 2009;50:612–621. doi: 10.1002/hep.23043. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Brown BD, Shi CX, Powell S, Hurlbut D, Graham FL, Lillicrap D. Helper-dependent adenoviral vectors mediate therapeutic factor VIII expression for several months with minimal accompanying toxicity in a canine model of severe hemophilia a. Blood. 2004;103:804–810. doi: 10.1182/blood-2003-05-1426. [DOI] [PubMed] [Google Scholar]
  28. Brunetti-Pierri N, Ng P. Helper-dependent adenoviral vectors for liver-directed gene therapy. Hum Mol Genet. 2011;20:R7–13. doi: 10.1093/hmg/ddr143. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Brunetti-Pierri N, Palmer DJ, Beaudet AL, Carey KD, Finegold M, Ng P. Acute toxicity after high-dose systemic injection of helper-dependent adenoviral vectors into nonhuman primates. Hum Gene Ther. 2004;15:35–46. doi: 10.1089/10430340460732445. [DOI] [PubMed] [Google Scholar]
  30. Brunetti-Pierri N, Nichols TC, McCorquodale S, et al. Sustained phenotypic correction of canine hemophilia B after systemic administration of helper-dependent adenoviral vector. Hum Gene Ther. 2005;16:811–820. doi: 10.1089/hum.2005.16.811. [DOI] [PubMed] [Google Scholar]
  31. Brunetti-Pierri N, Ng T, Iannitti DA, et al. Improved hepatic transduction, reduced systemic vector dissemination, and long-term transgene expression by delivering helper-dependent adenoviral vectors into the surgically isolated liver of nonhuman primates. Hum Gene Ther. 2006;17:391–404. doi: 10.1089/hum.2006.17.391. [DOI] [PubMed] [Google Scholar]
  32. Bryant LM, Christopher DM, Giles AR et al (2013) Lessons learned from the clinical development and market authorization of Glybera. Hum Gene Ther Clin Dev 24:55–64 [DOI] [PMC free article] [PubMed]
  33. Buchlis G, Podsakoff GM, Radu A, et al. Factor IX expression in skeletal muscle of a severe hemophilia B patient 10 years after AAV-mediated gene transfer. Blood. 2012;119:3038–3041. doi: 10.1182/blood-2011-09-382317. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Buning H, Schmidt M. Adeno-associated vector toxicity-to be or not to be? Mol Ther. 2015;23:1673–1675. doi: 10.1038/mt.2015.182. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Calcedo R, Wilson JM (2013) Humoral immune response to AAV. Front Immunol 4:341 [DOI] [PMC free article] [PubMed]
  36. Calcedo R, Vandenberghe LH, Gao G, Lin J, Wilson JM (2009) Worldwide epidemiology of neutralizing antibodies to adeno-associated viruses. J Infect Dis 199:381–390 [DOI] [PMC free article] [PubMed]
  37. Calcedo R, Morizono H, Wang L, et al. Adeno-associated virus antibody profiles in newborns, children, and adolescents. Clin Vaccine Immunol. 2011;18:1586–1588. doi: 10.1128/CVI.05107-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Cantore A, Ranzani M, Bartholomae CC, et al. Liver-directed lentiviral gene therapy in a dog model of hemophilia B. Sci Transl Med. 2015;7:277ra228. doi: 10.1126/scitranslmed.aaa1405. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Cartier N, Hacein-Bey-Abina S, Bartholomae CC, et al. Hematopoietic stem cell gene therapy with a lentiviral vector in X-linked adrenoleukodystrophy. Science. 2009;326:818–823. doi: 10.1126/science.1171242. [DOI] [PubMed] [Google Scholar]
  40. Cassiday L. Medical research: Gene-therapy reboot. Nature. 2014;509:651–653. doi: 10.1038/nj7502-651a. [DOI] [PubMed] [Google Scholar]
  41. Castello R, Borzone R, D’Aria S, Annunziata P, Piccolo P, Brunetti-Pierri N. Helper-dependent adenoviral vectors for liver-directed gene therapy of primary hyperoxaluria type 1. Gene Ther. 2016;23:129–134. doi: 10.1038/gt.2015.107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Cavazzana-Calvo M, Hacein-Bey S, de Saint BG, et al. Gene therapy of human severe combined immunodeficiency (SCID)-X1 disease. Science. 2000;288:669–672. doi: 10.1126/science.288.5466.669. [DOI] [PubMed] [Google Scholar]
  43. Cerreto M, Mehdawy B, Ombrone D, et al. Reversal of metabolic and neurological symptoms of phenylketonuric mice treated with a PAH containing helper-dependent adenoviral vector. Curr Gene Ther. 2012;12:48–56. doi: 10.2174/156652312799789280. [DOI] [PubMed] [Google Scholar]
  44. Chandler RJ, Venditti CP. Gene therapy for metabolic diseases. Transl Sci Rare Dis. 2016;1:73–89. doi: 10.3233/TRD-160007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Chandler RJ, LaFave MC, Varshney GK, et al. Vector design influences hepatic genotoxicity after adeno-associated virus gene therapy. J Clin Invest. 2015;125:870–880. doi: 10.1172/JCI79213. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Chandler RJ, Sands MS, Venditti CP. rAAV integration and genotoxicity: insights from animal models. Hum Gene Ther. 2017;28:314–322. doi: 10.1089/hum.2017.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Chang AH, Stephan MT, Lisowski L, Sadelain M. Erythroid-specific human factor IX delivery from in vivo selected hematopoietic stem cells following nonmyeloablative conditioning in hemophilia B mice. Mol Ther. 2008;16:1745–1752. doi: 10.1038/mt.2008.161. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Chen Y, Schroeder JA, Kuether EL, Zhang G, Shi Q. Platelet gene therapy by lentiviral gene delivery to hematopoietic stem cells restores hemostasis and induces humoral immune tolerance in FIX(null) mice. Mol Ther. 2014;22:169–177. doi: 10.1038/mt.2013.197. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Chicoine LG, Montgomery CL, Bremer WG, et al. Plasmapheresis eliminates the negative impact of AAV antibodies on microdystrophin gene expression following vascular delivery. Mol Ther. 2014;22:338–347. doi: 10.1038/mt.2013.244. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Chira S, Jackson CS, Oprea I, et al. Progresses towards safe and efficient gene therapy vectors. Oncotarget. 2015;6:30675–30703. doi: 10.18632/oncotarget.5169. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Choudhury SR, Fitzpatrick Z, Harris AF, et al. In vivo selection yields AAV-B1 capsid for central nervous system and muscle Gene therapy. Mol Ther. 2016;24:1247–1257. doi: 10.1038/mt.2016.84. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Chuah MK, Collen D, VandenDriessche T. Clinical gene transfer studies for hemophilia a. Semin Thromb Hemost. 2004;30:249–256. doi: 10.1055/s-2004-825638. [DOI] [PubMed] [Google Scholar]
  53. Cideciyan AV, Aleman TS, Boye SL et al (2008) Human gene therapy for RPE65 isomerase deficiency activates the retinoid cycle of vision but with slow rod kinetics. Proc Natl Acad Sci USA 105:15112–15117 [DOI] [PMC free article] [PubMed]
  54. Clayton PT. Inborn errors presenting with liver dysfunction. Semin Neonatol. 2002;7:49–63. doi: 10.1053/siny.2001.0086. [DOI] [PubMed] [Google Scholar]
  55. Clement N, Grieger JC (2016) Manufacturing of recombinant adeno-associated viral vectors for clinical trials. Mol Ther Methods Clin Dev 3:16002 [DOI] [PMC free article] [PubMed]
  56. Condiotti R, Goldenberg D, Giladi H, et al. Transduction of fetal mice with a feline lentiviral vector induces liver tumors which exhibit an E2F activation signature. Mol Ther. 2014;22:59–68. doi: 10.1038/mt.2013.193. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Cooper M, Nayak S, Hoffman BE, Terhorst C, Cao O, Herzog RW. Improved induction of immune tolerance to factor IX by hepatic AAV-8 gene transfer. Hum Gene Ther. 2009;20:767–776. doi: 10.1089/hum.2008.161. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Coppoletta JM, Wolbach SB. Body length and organ weights of infants and children: a study of the Body length and normal weights of the more important vital organs of the Body between birth and twelve years of age. Am J Pathol. 1933;9:55–70. [PMC free article] [PubMed] [Google Scholar]
  59. Corti M, Elder M, Falk D et al (2014) B-cell depletion is protective against anti-AAV capsid immune response: a human subject case study. Mol Ther Methods Clin Dev 1:14033 [DOI] [PMC free article] [PubMed]
  60. Counsell JR, Asgarian Z, Meng J, Ferrer V, Vink CA, Howe SJ, Waddington SN, Thrasher AJ, Muntoni F, Morgan JE, Danos O. Lentiviral vectors can be used for full-length dystrophin gene therapy. Sci Rep. 2017;7:79. doi: 10.1038/s41598-017-00152-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Couto L, Parker A, Gordon JW. Direct exposure of mouse spermatozoa to very high concentrations of a serotype-2 adeno-associated virus gene therapy vector fails to lead to germ cell transduction. Hum Gene Ther. 2004;15:287–291. doi: 10.1089/104303404322886138. [DOI] [PubMed] [Google Scholar]
  62. Crane B, Luo X, Demaster A, et al. Rescue administration of a helper-dependent adenovirus vector with long-term efficacy in dogs with glycogen storage disease type Ia. Gene Ther. 2012;19:443–452. doi: 10.1038/gt.2011.86. [DOI] [PubMed] [Google Scholar]
  63. Cunningham SC, Dane AP, Spinoulas A, Logan GJ, Alexander IE. Gene delivery to the juvenile mouse liver using AAV2/8 vectors. Mol Ther. 2008;16:1081–1088. doi: 10.1038/mt.2008.72. [DOI] [PubMed] [Google Scholar]
  64. Cunningham SC, Kok CY, Dane AP, et al. Induction and prevention of severe hyperammonemia in the spfash mouse model of ornithine transcarbamylase deficiency using shRNA and rAAV-mediated gene delivery. Mol Ther. 2011;19:854–859. doi: 10.1038/mt.2011.32. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Cunningham SC, Siew SM, Hallwirth CV, et al. Modeling correction of severe urea cycle defects in the growing murine liver using a hybrid recombinant adeno-associated virus/piggyBac transposase gene delivery system. Hepatology. 2015;62:417–428. doi: 10.1002/hep.27842. [DOI] [PubMed] [Google Scholar]
  66. D’Avola D, Lopez-Franco E, Sangro B, et al. Phase I open label liver-directed gene therapy clinical trial for acute intermittent porphyria. J Hepatol. 2016;65:776–783. doi: 10.1016/j.jhep.2016.05.012. [DOI] [PubMed] [Google Scholar]
  67. Deverman BE, Pravdo PL, Simpson BP, et al. Cre-dependent selection yields AAV variants for widespread gene transfer to the adult brain. Nat Biotechnol. 2016;34:204–209. doi: 10.1038/nbt.3440. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Dhawan A, Mitry RR, Hughes RD. Hepatocyte transplantation for liver-based metabolic disorders. J Inherit Metab Dis. 2006;29:431–435. doi: 10.1007/s10545-006-0245-8. [DOI] [PubMed] [Google Scholar]
  69. Doenecke A, Kromer A, Scherer MN, Schlitt HJ, Geissler EK. AAV plasmid DNA simplifies liver-directed in vivo gene therapy: comparison of expression levels after plasmid DNA-, adeno-associated virus- and adenovirus-mediated liver transfection. J Gene Med. 2010;12:810–817. doi: 10.1002/jgm.1498. [DOI] [PubMed] [Google Scholar]
  70. Dolgin E. Early clinical data raise the bar for hemophilia gene therapies. Nat Biotechnol. 2016;34:999–1001. doi: 10.1038/nbt1016-999. [DOI] [PubMed] [Google Scholar]
  71. Donsante A, Miller DG, Li Y, et al. AAV vector integration sites in mouse hepatocellular carcinoma. Science. 2007;317:477. doi: 10.1126/science.1142658. [DOI] [PubMed] [Google Scholar]
  72. Duan D, Yue Y, Engelhardt JF. Expanding AAV packaging capacity with trans-splicing or overlapping vectors: a quantitative comparison. Mol Ther. 2001;4:383–391. doi: 10.1006/mthe.2001.0456. [DOI] [PubMed] [Google Scholar]
  73. Earley LF, Powers JM, Adachi K, et al. Adeno-associated virus (AAV) assembly-activating protein is not an essential requirement for capsid assembly of AAV serotypes 4, 5, and 11. J Virol. 2017;91:e01980-16. doi: 10.1128/JVI.01980-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Ehrhardt A, Kay MA. A new adenoviral helper-dependent vector results in long-term therapeutic levels of human coagulation factor IX at low doses in vivo. Blood. 2002;99:3923–3930. doi: 10.1182/blood.v99.11.3923. [DOI] [PubMed] [Google Scholar]
  75. Ehrhardt A, Xu H, Dillow AM, Bellinger DA, Nichols TC, Kay MA. A gene-deleted adenoviral vector results in phenotypic correction of canine hemophilia B without liver toxicity or thrombocytopenia. Blood. 2003;102:2403–2411. doi: 10.1182/blood-2003-01-0314. [DOI] [PubMed] [Google Scholar]
  76. Elsabahy M, Nazarali A, Foldvari M. Non-viral nucleic acid delivery: key challenges and future directions. Curr Drug Deliv. 2011;8:235–244. doi: 10.2174/156720111795256174. [DOI] [PubMed] [Google Scholar]
  77. Fahs SA, Hille MT, Shi Q, Weiler H, Montgomery RR. A conditional knockout mouse model reveals endothelial cells as the principal and possibly exclusive source of plasma factor VIII. Blood. 2014;123:3706–3713. doi: 10.1182/blood-2014-02-555151. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Fischer A, Hacein-Bey-Abina S, Cavazzana-Calvo M. 20 years of gene therapy for SCID. Nat Immunol. 2010;11:457–460. doi: 10.1038/ni0610-457. [DOI] [PubMed] [Google Scholar]
  79. Flanigan KM, Campbell K, Viollet L, et al. Anti-dystrophin T cell responses in Duchenne muscular dystrophy: prevalence and a glucocorticoid treatment effect. Hum Gene Ther. 2013;24:797–806. doi: 10.1089/hum.2013.092. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Flotte TR, Brantly ML, Spencer LT, et al. Phase I trial of intramuscular injection of a recombinant adeno-associated virus alpha 1-antitrypsin (rAAV2-CB-hAAT) gene vector to AAT-deficient adults. Hum Gene Ther. 2004;15:93–128. doi: 10.1089/10430340460732490. [DOI] [PubMed] [Google Scholar]
  81. Flotte TR, Trapnell BC, Humphries M, et al. Phase 2 clinical trial of a recombinant adeno-associated viral vector expressing alpha1-antitrypsin: interim results. Hum Gene Ther. 2011;22:1239–1247. doi: 10.1089/hum.2011.053. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Fumoto S, Kawakami S, Hashida M, Nishida K (2013) Targeted gene delivery: importance of administration routes. In: Good D, Wei M (eds.) Novel gene therapy approaches. doi: 10.5772/54741
  83. Gao GP, Yang Y, Wilson JM. Biology of adenovirus vectors with E1 and E4 deletions for liver-directed gene therapy. J Virol. 1996;70:8934–8943. doi: 10.1128/jvi.70.12.8934-8943.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Gebhardt R, Matz-Soja M. Liver zonation: novel aspects of its regulation and its impact on homeostasis. World J Gastroenterol. 2014;20:8491–8504. doi: 10.3748/wjg.v20.i26.8491. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. George L, Sullivan S, Giermasz A (2016) Adeno-associated virus mediated gene transfer for Hemophilia B achieves sustained mean factor IX activity levels of >30% without immunosuppression. 58th annual meeting of the American Society of Hemophilia
  86. Gil-Farina I, Schmidt M. Interaction of vectors and parental viruses with the host genome. Curr Opin Virol. 2016;21:35–40. doi: 10.1016/j.coviro.2016.07.004. [DOI] [PubMed] [Google Scholar]
  87. Ginn SL, Alexander IE, Edelstein ML, Abedi MR, Wixon J. Gene therapy clinical trials worldwide to 2012 - an update. J Gene Med. 2013;15:65–77. doi: 10.1002/jgm.2698. [DOI] [PubMed] [Google Scholar]
  88. Gramignoli R, Vosough M, Kannisto K, Srinivasan RC, Strom SC. Clinical hepatocyte transplantation: practical limits and possible solutions. Eur Surg Res. 2015;54:162–177. doi: 10.1159/000369552. [DOI] [PubMed] [Google Scholar]
  89. Grieger JC, Soltys SM, Samulski RJ. Production of recombinant adeno-associated virus vectors using suspension HEK293 cells and continuous harvest of vector from the culture media for GMP FIX and FLT1 clinical vector. Mol Ther. 2016;24:287–297. doi: 10.1038/mt.2015.187. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Grossman M, Raper SE, Kozarsky K, et al. Successful ex vivo gene therapy directed to liver in a patient with familial hypercholesterolaemia. Nat Genet. 1994;6:335–341. doi: 10.1038/ng0494-335. [DOI] [PubMed] [Google Scholar]
  91. Grossman M, Rader DJ, Muller DW, et al. A pilot study of ex vivo gene therapy for homozygous familial hypercholesterolaemia. Nat Med. 1995;1:1148–1154. doi: 10.1038/nm1195-1148. [DOI] [PubMed] [Google Scholar]
  92. Haberle J, Boddaert N, Burlina A, et al. Suggested guidelines for the diagnosis and management of urea cycle disorders. Orphanet Journal of Rare Diseases. 2012;7:32. doi: 10.1186/1750-1172-7-32. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Hacein-Bey-Abina S, von Kalle C, Schmidt M, et al. A serious adverse event after successful gene therapy for X-linked severe combined immunodeficiency. N Engl J Med. 2003;348:255–256. doi: 10.1056/NEJM200301163480314. [DOI] [PubMed] [Google Scholar]
  94. Hacein-Bey-Abina S, Von Kalle C, Schmidt M, et al. LMO2-associated clonal T cell proliferation in two patients after gene therapy for SCID-X1. Science. 2003;302:415–419. doi: 10.1126/science.1088547. [DOI] [PubMed] [Google Scholar]
  95. Hastie E, Samulski RJ. Recombinant adeno-associated virus vectors in the treatment of rare diseases. Expert Opin Orphan Drugs. 2015;3:675–689. doi: 10.1517/21678707.2015.1039511. [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Herzog RW, Hagstrom JN, Kung SH et al (1997) Stable gene transfer and expression of human blood coagulation factor IX after intramuscular injection of recombinant adeno-associated virus. Proc Natl Acad Sci USA 94:5804–5809 [DOI] [PMC free article] [PubMed]
  97. Herzog RW, Yang EY, Couto LB, et al. Long-term correction of canine hemophilia B by gene transfer of blood coagulation factor IX mediated by adeno-associated viral vector. Nat Med. 1999;5:56–63. doi: 10.1038/4743. [DOI] [PubMed] [Google Scholar]
  98. Hickey RD, Mao SA, Glorioso J, et al. Curative ex vivo liver-directed gene therapy in a pig model of hereditary tyrosinemia type 1. Sci Transl Med. 2016;8:349ra399. doi: 10.1126/scitranslmed.aaf3838. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Hirsch ML, Wolf SJ, Samulski RJ. Delivering transgenic DNA exceeding the carrying capacity of AAV vectors. Methods Mol Biol. 2016;1382:21–39. doi: 10.1007/978-1-4939-3271-9_2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Hoggatt J. Gene therapy for "bubble boy" disease. Cell. 2016;166:263. doi: 10.1016/j.cell.2016.06.049. [DOI] [PubMed] [Google Scholar]
  101. Hou X, Jiao R, Guo X, et al. Construction of minicircle DNA vectors capable of correcting familial hypercholesterolemia phenotype in a LDLR-deficient mouse model. Gene Ther. 2016;23:657–663. doi: 10.1038/gt.2016.37. [DOI] [PubMed] [Google Scholar]
  102. Hu C, Cela RG, Suzuki M, Lee B, Lipshutz GS (2011) Neonatal helper-dependent adenoviral vector gene therapy mediates correction of hemophilia a and tolerance to human factor VIII. Proc Natl Acad Sci USA 108:2082–2087 [DOI] [PMC free article] [PubMed]
  103. Jayandharan GR, Aslanidi G, Martino AT et al (2011) Activation of the NF-kappaB pathway by adeno-associated virus (AAV) vectors and its implications in immune response and gene therapy. Proc Natl Acad Sci USA 108:3743–3748 [DOI] [PMC free article] [PubMed] [Retracted]
  104. Jiang H, Couto LB, Patarroyo-White S, et al. Effects of transient immunosuppression on adenoassociated, virus-mediated, liver-directed gene transfer in rhesus macaques and implications for human gene therapy. Blood. 2006;108:3321–3328. doi: 10.1182/blood-2006-04-017913. [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Jiang H, Pierce GF, Ozelo MC, et al. Evidence of multiyear factor IX expression by AAV-mediated gene transfer to skeletal muscle in an individual with severe hemophilia B. Mol Ther. 2006;14:452–455. doi: 10.1016/j.ymthe.2006.05.004. [DOI] [PubMed] [Google Scholar]
  106. Jooss K, Chirmule N. Immunity to adenovirus and adeno-associated viral vectors: implications for gene therapy. Gene Ther. 2003;10:955–963. doi: 10.1038/sj.gt.3302037. [DOI] [PubMed] [Google Scholar]
  107. Junge N, Mingozzi F, Ott M, Baumann U. Adeno-associated virus vector-based gene therapy for monogenetic metabolic diseases of the liver. J Pediatr Gastroenterol Nutr. 2015;60:433–440. doi: 10.1097/MPG.0000000000000703. [DOI] [PubMed] [Google Scholar]
  108. Kaeppel C, Beattie SG, Fronza R, et al. A largely random AAV integration profile after LPLD gene therapy. Nat Med. 2013;19:889–891. doi: 10.1038/nm.3230. [DOI] [PubMed] [Google Scholar]
  109. Kalyuzhniy O, Di Paolo NC, Silvestry M et al (2008) Adenovirus serotype 5 hexon is critical for virus infection of hepatocytes in vivo. Proc Natl Acad Sci USA 105:5483–5488 [DOI] [PMC free article] [PubMed]
  110. Kattenhorn LM, Tipper CH, Stoica L, et al. Adeno-associated virus Gene therapy for liver disease. Hum Gene Ther. 2016;27:947–961. doi: 10.1089/hum.2016.160. [DOI] [PMC free article] [PubMed] [Google Scholar]
  111. Kay MA. State-of-the-art gene-based therapies: the road ahead. Nat Rev Genet. 2011;12:316–328. doi: 10.1038/nrg2971. [DOI] [PubMed] [Google Scholar]
  112. Kay MA, Manno CS, Ragni MV, et al. Evidence for gene transfer and expression of factor IX in haemophilia B patients treated with an AAV vector. Nat Genet. 2000;24:257–261. doi: 10.1038/73464. [DOI] [PubMed] [Google Scholar]
  113. Kay MA, He CY, Chen ZY. A robust system for production of minicircle DNA vectors. Nat Biotechnol. 2010;28:1287–1289. doi: 10.1038/nbt.1708. [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Kiang A, Hartman ZC, Everett RS, et al. Multiple innate inflammatory responses induced after systemic adenovirus vector delivery depend on a functional complement system. Mol Ther. 2006;14:588–598. doi: 10.1016/j.ymthe.2006.03.024. [DOI] [PubMed] [Google Scholar]
  115. Kim IH, Jozkowicz A, Piedra PA, Oka K, Chan L (2001) Lifetime correction of genetic deficiency in mice with a single injection of helper-dependent adenoviral vector. Proc Natl Acad Sci USA 98:13282–13287 [DOI] [PMC free article] [PubMed]
  116. Koeberl DD, Sun B, Bird A, Chen YT, Oka K, Chan L. Efficacy of helper-dependent adenovirus vector-mediated gene therapy in murine glycogen storage disease type Ia. Mol Ther. 2007;15:1253–1258. doi: 10.1038/sj.mt.6300188. [DOI] [PubMed] [Google Scholar]
  117. Koo T, Popplewell L, Athanasopoulos T, Dickson G. Triple trans-splicing adeno-associated virus vectors capable of transferring the coding sequence for full-length dystrophin protein into dystrophic mice. Hum Gene Ther. 2014;25:98–108. doi: 10.1089/hum.2013.164. [DOI] [PubMed] [Google Scholar]
  118. Kotterman MA, Schaffer DV. Engineering adeno-associated viruses for clinical gene therapy. Nat Rev Genet. 2014;15:445–451. doi: 10.1038/nrg3742. [DOI] [PMC free article] [PubMed] [Google Scholar]
  119. Kotterman MA, Chalberg TW, Schaffer DV (2015) Viral vectors for gene therapy: translational and clinical outlook. Annu Rev Biomed Eng 17:63–89 [DOI] [PubMed]
  120. Kuether EL, Schroeder JA, Fahs SA, et al. Lentivirus-mediated platelet gene therapy of murine hemophilia a with pre-existing anti-factor VIII immunity. J Thromb Haemost. 2012;10:1570–1580. doi: 10.1111/j.1538-7836.2012.04791.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  121. Ledley FD, McNamee LM, Uzdil V, Morgan IW. Why commercialization of gene therapy stalled; examining the life cycles of gene therapy technologies. Gene Ther. 2014;21:188–194. doi: 10.1038/gt.2013.72. [DOI] [PubMed] [Google Scholar]
  122. Li C, Narkbunnam N, Samulski RJ, et al. Neutralizing antibodies against adeno-associated virus examined prospectively in pediatric patients with hemophilia. Gene Ther. 2012;19:288–294. doi: 10.1038/gt.2011.90. [DOI] [PubMed] [Google Scholar]
  123. Li M, Dick A, Montenovo M, Horslen S, Hansen R. Cost-effectiveness of liver transplantation in methylmalonic and propionic acidemias. Liver Transpl. 2015;21:1208–1218. doi: 10.1002/lt.24173. [DOI] [PubMed] [Google Scholar]
  124. Lipshutz GS, Titre D, Brindle M, Bisconte AR, Contag CH, Gaensler KM. Comparison of gene expression after intraperitoneal delivery of AAV2 or AAV5 in utero. Mol Ther. 2003;8:90–98. doi: 10.1016/s1525-0016(03)00132-1. [DOI] [PubMed] [Google Scholar]
  125. Lisowski L, Dane AP, Chu K, et al. Selection and evaluation of clinically relevant AAV variants in a xenograft liver model. Nature. 2014;506:382–386. doi: 10.1038/nature12875. [DOI] [PMC free article] [PubMed] [Google Scholar]
  126. Liu F, Song Y, Liu D. Hydrodynamics-based transfection in animals by systemic administration of plasmid DNA. Gene Ther. 1999;6:1258–1266. doi: 10.1038/sj.gt.3300947. [DOI] [PubMed] [Google Scholar]
  127. Louis Jeune V, Joergensen JA, Hajjar RJ, Weber T. Pre-existing anti-adeno-associated virus antibodies as a challenge in AAV gene therapy. Hum Gene Ther Methods. 2013;24:59–67. doi: 10.1089/hgtb.2012.243. [DOI] [PMC free article] [PubMed] [Google Scholar]
  128. Lozier JN, Metzger ME, Donahue RE, Morgan RA. Adenovirus-mediated expression of human coagulation factor IX in the rhesus macaque is associated with dose-limiting toxicity. Blood. 1999;94:3968–3975. [PubMed] [Google Scholar]
  129. Lu J, Williams JA, Luke J, Zhang F, Chu K, Kay MA. A 5′ noncoding exon containing engineered intron enhances transgene expression from recombinant AAV vectors in vivo. Hum Gene Ther. 2017;28:125–134. doi: 10.1089/hum.2016.140. [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Macaulay R. How ready are european payers for ema adaptive pathways? Value Health. 2015;18:A341. [Google Scholar]
  131. Maguire AM, Simonelli F, Pierce EA, et al. Safety and efficacy of gene transfer for Leber’s congenital amaurosis. N Engl J Med. 2008;358:2240–2248. doi: 10.1056/NEJMoa0802315. [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Mah C, Sarkar R, Zolotukhin I, et al. Dual vectors expressing murine factor VIII result in sustained correction of hemophilia a mice. Hum Gene Ther. 2003;14:143–152. doi: 10.1089/104303403321070838. [DOI] [PubMed] [Google Scholar]
  133. Majhen D, Calderon H, Chandra N, et al. Adenovirus-based vaccines for fighting infectious diseases and cancer: progress in the field. Hum Gene Ther. 2014;25:301–317. doi: 10.1089/hum.2013.235. [DOI] [PubMed] [Google Scholar]
  134. Manno CS, Chew AJ, Hutchison S, et al. AAV-mediated factor IX gene transfer to skeletal muscle in patients with severe hemophilia B. Blood. 2003;101:2963–2972. doi: 10.1182/blood-2002-10-3296. [DOI] [PubMed] [Google Scholar]
  135. Manno CS, Pierce GF, Arruda VR, et al. Successful transduction of liver in hemophilia by AAV-factor IX and limitations imposed by the host immune response. Nat Med. 2006;12:342–347. doi: 10.1038/nm1358. [DOI] [PubMed] [Google Scholar]
  136. Mattar CN, Nathwani AC, Waddington SN, et al. Stable human FIX expression after 0.9G intrauterine gene transfer of self-complementary adeno-associated viral vector 5 and 8 in macaques. Mol Ther. 2011;19:1950–1960. doi: 10.1038/mt.2011.107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  137. Mayrhofer P, Schleef M, Jechlinger W. Use of minicircle plasmids for gene therapy. Methods Mol Biol. 2009;542:87–104. doi: 10.1007/978-1-59745-561-9_4. [DOI] [PubMed] [Google Scholar]
  138. McCarty DM. Self-complementary AAV vectors; advances and applications. Mol Ther. 2008;16:1648–1656. doi: 10.1038/mt.2008.171. [DOI] [PubMed] [Google Scholar]
  139. McCarty DM, Young SM, Jr, Samulski RJ. Integration of adeno-associated virus (AAV) and recombinant AAV vectors. Annu Rev Genet. 2004;38:819–845. doi: 10.1146/annurev.genet.37.110801.143717. [DOI] [PubMed] [Google Scholar]
  140. McCormack WM, Jr, Seiler MP, Bertin TK, et al. Helper-dependent adenoviral gene therapy mediates long-term correction of the clotting defect in the canine hemophilia a model. J Thromb Haemost. 2006;4:1218–1225. doi: 10.1111/j.1538-7836.2006.01901.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  141. McGarrity GJ, Hoyah G, Winemiller A, et al. Patient monitoring and follow-up in lentiviral clinical trials. J Gene Med. 2013;15:78–82. doi: 10.1002/jgm.2691. [DOI] [PubMed] [Google Scholar]
  142. McIntosh JH, Cochrane M, Cobbold S, et al. Successful attenuation of humoral immunity to viral capsid and transgenic protein following AAV-mediated gene transfer with a non-depleting CD4 antibody and cyclosporine. Gene Ther. 2012;19:78–85. doi: 10.1038/gt.2011.64. [DOI] [PMC free article] [PubMed] [Google Scholar]
  143. McKay TR, Rahim AA, Buckley SM, et al. Perinatal gene transfer to the liver. Curr Pharm Des. 2011;17:2528–2541. doi: 10.2174/138161211797247541. [DOI] [PMC free article] [PubMed] [Google Scholar]
  144. Mian A, Lee B. Urea-cycle disorders as a paradigm for inborn errors of hepatocyte metabolism. Trends Mol Med. 2002;8:583–589. doi: 10.1016/s1471-4914(02)02437-1. [DOI] [PubMed] [Google Scholar]
  145. Miesbach W, Tangelder M, Klamroth R. Updated results from a dose escalating study in adult patients with haemophilia B treated with AMT-060 (AAV5-hFIX) gene therapy. Haemophilia. 2016;22:151. [Google Scholar]
  146. Miller DG, Adam MA, Miller AD. Gene transfer by retrovirus vectors occurs only in cells that are actively replicating at the time of infection. Mol Cell Biol. 1990;10:4239–4242. doi: 10.1128/mcb.10.8.4239. [DOI] [PMC free article] [PubMed] [Google Scholar]
  147. Mingozzi F, High KA. Immune responses to AAV in clinical trials. Curr Gene Ther. 2007;7:316–324. doi: 10.2174/156652307782151425. [DOI] [PubMed] [Google Scholar]
  148. Mingozzi F, High KA. Immune responses to AAV in clinical trials. Curr Gene Ther. 2011;11:321–330. doi: 10.2174/156652311796150354. [DOI] [PubMed] [Google Scholar]
  149. Mingozzi F, High KA. Immune responses to AAV vectors: overcoming barriers to successful gene therapy. Blood. 2013;122:23–36. doi: 10.1182/blood-2013-01-306647. [DOI] [PMC free article] [PubMed] [Google Scholar]
  150. Mingozzi F, Maus MV, Hui DJ, et al. CD8(+) T-cell responses to adeno-associated virus capsid in humans. Nat Med. 2007;13:419–422. doi: 10.1038/nm1549. [DOI] [PubMed] [Google Scholar]
  151. Mingozzi F, Chen Y, Murphy SL, et al. Pharmacological modulation of humoral immunity in a nonhuman primate model of AAV gene transfer for hemophilia B. Mol Ther. 2012;20:1410–1416. doi: 10.1038/mt.2012.84. [DOI] [PMC free article] [PubMed] [Google Scholar]
  152. Mingozzi F, Chen Y, Edmonson SC, et al. Prevalence and pharmacological modulation of humoral immunity to AAV vectors in gene transfer to synovial tissue. Gene Ther. 2013;20:417–424. doi: 10.1038/gt.2012.55. [DOI] [PMC free article] [PubMed] [Google Scholar]
  153. Miyoshi H, Blomer U, Takahashi M, Gage FH, Verma IM. Development of a self-inactivating lentivirus vector. J Virol. 1998;72:8150–8157. doi: 10.1128/jvi.72.10.8150-8157.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  154. Monahan PE, Lothrop CD, Sun J, et al. Proteasome inhibitors enhance gene delivery by AAV virus vectors expressing large genomes in hemophilia mouse and dog models: a strategy for broad clinical application. Mol Ther. 2010;18:1907–1916. doi: 10.1038/mt.2010.170. [DOI] [PMC free article] [PubMed] [Google Scholar]
  155. Monahan P, Walsh C, Powell J. Update on a phase 1/2 open-label trial of BAX335, an adeno-associated virus 8 (AAV8) vector-based gene therapy program for hemophilia B. J Thromb Haemost. 2015;13(Suppl 2):87. [Google Scholar]
  156. Monteilhet V, Saheb S, Boutin S, et al. A 10 patient case report on the impact of plasmapheresis upon neutralizing factors against adeno-associated virus (AAV) types 1, 2, 6, and 8. Mol Ther. 2011;19:2084–2091. doi: 10.1038/mt.2011.108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  157. Morral N, O’Neal W, Rice K et al (1999) Administration of helper-dependent adenoviral vectors and sequential delivery of different vector serotype for long-term liver-directed gene transfer in baboons. Proc Natl Acad Sci USA 96:12816–12821 [DOI] [PMC free article] [PubMed]
  158. Morral N, O’Neal WK, Rice K, et al. Lethal toxicity, severe endothelial injury, and a threshold effect with high doses of an adenoviral vector in baboons. Hum Gene Ther. 2002;13:143–154. doi: 10.1089/10430340152712692. [DOI] [PubMed] [Google Scholar]
  159. Mukherjee S, Thrasher AJ. Gene therapy for PIDs: progress, pitfalls and prospects. Gene. 2013;525:174–181. doi: 10.1016/j.gene.2013.03.098. [DOI] [PMC free article] [PubMed] [Google Scholar]
  160. Mullard A. European regulators launch "breakthrough" equivalent programme. Nat Rev Drug Discov. 2016;15:223. [Google Scholar]
  161. Muruve DA, Cotter MJ, Zaiss AK, et al. Helper-dependent adenovirus vectors elicit intact innate but attenuated adaptive host immune responses in vivo. J Virol. 2004;78:5966–5972. doi: 10.1128/JVI.78.11.5966-5972.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  162. Narayanan G, Cossu G, Galli MC et al (2014) Clinical development of gene therapy needs a tailored approach: a regulatory perspective from the European Union. Hum Gene Ther Clin Dev 25:1–6 [DOI] [PubMed]
  163. Nathwani AC, Gray JT, McIntosh J, et al. Safe and efficient transduction of the liver after peripheral vein infusion of self-complementary AAV vector results in stable therapeutic expression of human FIX in nonhuman primates. Blood. 2007;109:1414–1421. doi: 10.1182/blood-2006-03-010181. [DOI] [PMC free article] [PubMed] [Google Scholar]
  164. Nathwani AC, Rosales C, McIntosh J, et al. Long-term safety and efficacy following systemic administration of a self-complementary AAV vector encoding human FIX pseudotyped with serotype 5 and 8 capsid proteins. Mol Ther. 2011;19:876–885. doi: 10.1038/mt.2010.274. [DOI] [PMC free article] [PubMed] [Google Scholar]
  165. Nathwani AC, Reiss UM, Tuddenham EG, et al. Long-term safety and efficacy of factor IX gene therapy in hemophilia B. N Engl J Med. 2014;371:1994–2004. doi: 10.1056/NEJMoa1407309. [DOI] [PMC free article] [PubMed] [Google Scholar]
  166. Nault JC, Datta S, Imbeaud S, et al. Recurrent AAV2-related insertional mutagenesis in human hepatocellular carcinomas. Nat Genet. 2015;47:1187–1193. doi: 10.1038/ng.3389. [DOI] [PubMed] [Google Scholar]
  167. Naumer M, Sonntag F, Schmidt K, et al. Properties of the adeno-associated virus assembly-activating protein. J Virol. 2012;86:13038–13048. doi: 10.1128/JVI.01675-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  168. Nguyen TH, Mainot S, Lainas P, et al. Ex vivo liver-directed gene therapy for the treatment of metabolic diseases: advances in hepatocyte transplantation and retroviral vectors. Curr Gene Ther. 2009;9:136–149. doi: 10.2174/156652309787909481. [DOI] [PubMed] [Google Scholar]
  169. Nienhuis AW, Nathwani AC, Davidoff AM. Gene therapy for hemophilia. Hum Gene Ther. 2016;27:305–308. doi: 10.1089/hum.2016.018. [DOI] [PubMed] [Google Scholar]
  170. Nightingale SJ, Hollis RP, Pepper KA, et al. Transient gene expression by nonintegrating lentiviral vectors. Mol Ther. 2006;13:1121–1132. doi: 10.1016/j.ymthe.2006.01.008. [DOI] [PubMed] [Google Scholar]
  171. Nomura S, Merched A, Nour E, Dieker C, Oka K, Chan L. Low-density lipoprotein receptor gene therapy using helper-dependent adenovirus produces long-term protection against atherosclerosis in a mouse model of familial hypercholesterolemia. Gene Ther. 2004;11:1540–1548. doi: 10.1038/sj.gt.3302310. [DOI] [PubMed] [Google Scholar]
  172. Nowrouzi A, Cheung WT, Li T, et al. The fetal mouse is a sensitive genotoxicity model that exposes lentiviral-associated mutagenesis resulting in liver oncogenesis. Mol Ther. 2013;21:324–337. doi: 10.1038/mt.2012.224. [DOI] [PMC free article] [PubMed] [Google Scholar]
  173. Oishi Y, Kakimoto T, Yuan W, Kuno S, Yamashita H, Chiba T. Fetal Gene therapy for ornithine Transcarbamylase deficiency by intrahepatic plasmid DNA-micro-bubble injection combined with hepatic ultrasound Insonation. Ultrasound Med Biol. 2016;42:1357–1361. doi: 10.1016/j.ultrasmedbio.2015.10.007. [DOI] [PubMed] [Google Scholar]
  174. Orkin SH, Motulsky AG (1995) Report and recommendations of the panel to assess the NIH investment in research on gene therapy. National Institute of Health, Bethesda
  175. Orkin SH, Reilly P. MEDICINE. Paying for future success in gene therapy. Science. 2016;352:1059–1061. doi: 10.1126/science.aaf4770. [DOI] [PubMed] [Google Scholar]
  176. Pasi J, Wong W, Rangarajan S. Interim results of an open-label, phase 1/2 study of BMN 270, an AAV5-FVIII gene transfer in severe hemophilia a. Haemophilia. 2016;22:151. [Google Scholar]
  177. Perrin GQ, Zolotukhin I, Sherman A et al (2016) Dynamics of antigen presentation to transgene product-specific CD4+ T cells and of Treg induction upon hepatic AAV gene transfer. Mol Ther Methods Clin Dev 3:16083 [DOI] [PMC free article] [PubMed]
  178. Philippe S, Sarkis C, Barkats M et al (2006) Lentiviral vectors with a defective integrase allow efficient and sustained transgene expression in vitro and in vivo. Proc Natl Acad Sci USA 103:17684–17689 [DOI] [PMC free article] [PubMed]
  179. Piccolo P, Brunetti-Pierri N. Challenges and prospects for helper-dependent adenoviral vector-mediated gene therapy. Biomedicine. 2014;2:132–148. doi: 10.3390/biomedicines2020132. [DOI] [PMC free article] [PubMed] [Google Scholar]
  180. Pien GC, Basner-Tschakarjan E, Hui DJ, et al. Capsid antigen presentation flags human hepatocytes for destruction after transduction by adeno-associated viral vectors. J Clin Invest. 2009;119:1688–1695. doi: 10.1172/JCI36891. [DOI] [PMC free article] [PubMed] [Google Scholar]
  181. Pillay S, Meyer NL, Puschnik AS, et al. An essential receptor for adeno-associated virus infection. Nature. 2016;530:108–112. doi: 10.1038/nature16465. [DOI] [PMC free article] [PubMed] [Google Scholar]
  182. Powell JS, Ragni MV, White GC, 2nd, et al. Phase 1 trial of FVIII gene transfer for severe hemophilia a using a retroviral construct administered by peripheral intravenous infusion. Blood. 2003;102:2038–2045. doi: 10.1182/blood-2003-01-0167. [DOI] [PubMed] [Google Scholar]
  183. Puumalainen AM, Vapalahti M, Agrawal RS, et al. Beta-galactosidase gene transfer to human malignant glioma in vivo using replication-deficient retroviruses and adenoviruses. Hum Gene Ther. 1998;9:1769–1774. doi: 10.1089/hum.1998.9.12-1769. [DOI] [PubMed] [Google Scholar]
  184. Rabinowitz JE, Rolling F, Li C, et al. Cross-packaging of a single adeno-associated virus (AAV) type 2 vector genome into multiple AAV serotypes enables transduction with broad specificity. J Virol. 2002;76:791–801. doi: 10.1128/JVI.76.2.791-801.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  185. Rahim AA, Wong AM, Howe SJ, et al. Efficient gene delivery to the adult and fetal CNS using pseudotyped non-integrating lentiviral vectors. Gene Ther. 2009;16:509–520. doi: 10.1038/gt.2008.186. [DOI] [PubMed] [Google Scholar]
  186. Raper SE, Grossman M, Rader DJ, et al. Safety and feasibility of liver-directed ex vivo gene therapy for homozygous familial hypercholesterolemia. Ann Surg. 1996;223:116–126. doi: 10.1097/00000658-199602000-00002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  187. Raper SE, Haskal ZJ, Ye X, et al. Selective gene transfer into the liver of non-human primates with E1-deleted, E2A-defective, or E1-E4 deleted recombinant adenoviruses. Hum Gene Ther. 1998;9:671–679. doi: 10.1089/hum.1998.9.5-671. [DOI] [PubMed] [Google Scholar]
  188. Raper SE, Yudkoff M, Chirmule N, et al. A pilot study of in vivo liver-directed gene transfer with an adenoviral vector in partial ornithine transcarbamylase deficiency. Hum Gene Ther. 2002;13:163–175. doi: 10.1089/10430340152712719. [DOI] [PubMed] [Google Scholar]
  189. Raper SE, Chirmule N, Lee FS, et al. Fatal systemic inflammatory response syndrome in a ornithine transcarbamylase deficient patient following adenoviral gene transfer. Mol Genet Metab. 2003;80:148–158. doi: 10.1016/j.ymgme.2003.08.016. [DOI] [PubMed] [Google Scholar]
  190. Rastall DP, Seregin SS, Aldhamen YA, et al. Long-term, high-level hepatic secretion of acid alpha-glucosidase for Pompe disease achieved in non-human primates using helper-dependent adenovirus. Gene Ther. 2016;23:743–752. doi: 10.1038/gt.2016.53. [DOI] [PubMed] [Google Scholar]
  191. Reddy PS, Sakhuja K, Ganesh S, et al. Sustained human factor VIII expression in hemophilia a mice following systemic delivery of a gutless adenoviral vector. Mol Ther. 2002;5:63–73. doi: 10.1006/mthe.2001.0510. [DOI] [PubMed] [Google Scholar]
  192. Rosewell Shaw A, Suzuki M. Recent advances in oncolytic adenovirus therapies for cancer. Curr Opin Virol. 2016;21:9–15. doi: 10.1016/j.coviro.2016.06.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  193. Samulski RJ, Muzyczka N (2014) AAV-mediated gene therapy for research and therapeutic purposes. Annu Rev Virol 1:427–451 [DOI] [PubMed]
  194. Sarkar R, Mucci M, Addya S, et al. Long-term efficacy of adeno-associated virus serotypes 8 and 9 in hemophilia a dogs and mice. Hum Gene Ther. 2006;17:427–439. doi: 10.1089/hum.2006.17.427. [DOI] [PubMed] [Google Scholar]
  195. Schiedner G, Bloch W, Hertel S, et al. A hemodynamic response to intravenous adenovirus vector particles is caused by systemic Kupffer cell-mediated activation of endothelial cells. Hum Gene Ther. 2003;14:1631–1641. doi: 10.1089/104303403322542275. [DOI] [PubMed] [Google Scholar]
  196. Schimmer J, Breazzano S (2016) Investor outlook: solving gene therapy pricing...with a cures voucher? Hum Gene Ther Clin Dev 27:132–136 [DOI] [PubMed]
  197. Schmitt F, Pastore N, Abarrategui-Pontes C, et al. Correction of hyperbilirubinemia in gunn rats by surgical delivery of low doses of helper-dependent adenoviral vectors. Hum Gene Ther Methods. 2014;25:181–186. doi: 10.1089/hgtb.2013.236. [DOI] [PubMed] [Google Scholar]
  198. Sendra L, Miguel A, Perez-Enguix D, et al. Studying closed hydrodynamic models of "in vivo" DNA perfusion in pig liver for Gene therapy translation to humans. PLoS One. 2016;11:e0163898. doi: 10.1371/journal.pone.0163898. [DOI] [PMC free article] [PubMed] [Google Scholar]
  199. Shi Y, Falahati R, Zhang J, Flebbe-Rehwaldt L, Gaensler KM. Role of antigen-specific regulatory CD4+CD25+ T cells in tolerance induction after neonatal IP administration of AAV-hF.IX. Gene Ther. 2013;20:987–996. doi: 10.1038/gt.2013.22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  200. Sokal EM. Liver transplantation for inborn errors of liver metabolism. J Inherit Metab Dis. 2006;29:426–430. doi: 10.1007/s10545-006-0288-x. [DOI] [PubMed] [Google Scholar]
  201. Spada M, Riva S, Maggiore G, Cintorino D, Gridelli B. Pediatric liver transplantation. World J Gastroenterol. 2009;15:648–674. doi: 10.3748/wjg.15.648. [DOI] [PMC free article] [PubMed] [Google Scholar]
  202. Stephen SL, Montini E, Sivanandam VG, et al. Chromosomal integration of adenoviral vector DNA in vivo. J Virol. 2010;84:9987–9994. doi: 10.1128/JVI.00751-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  203. Summerford C, Samulski RJ. AAVR: a multi-serotype receptor for AAV. Mol Ther. 2016;24:663–666. doi: 10.1038/mt.2016.49. [DOI] [PMC free article] [PubMed] [Google Scholar]
  204. Sunderman FW, Boerner F (1949) Normal value in clinical medicine. Saunders, Philadelphia
  205. Suwanmanee T, Hu G, Gui T, et al. Integration-deficient lentiviral vectors expressing codon-optimized R338L human FIX restore normal hemostasis in hemophilia B mice. Mol Ther. 2014;22:567–574. doi: 10.1038/mt.2013.188. [DOI] [PMC free article] [PubMed] [Google Scholar]
  206. Suzuki T, Sasaki T, Yano K, et al. Development of a recombinant adenovirus vector production system free of replication-competent adenovirus by utilizing a packaging size limit of the viral genome. Virus Res. 2011;158:154–160. doi: 10.1016/j.virusres.2011.03.026. [DOI] [PubMed] [Google Scholar]
  207. Suzuki M, Bertin TK, Rogers GL, et al. Differential type I interferon-dependent transgene silencing of helper-dependent adenoviral vs. adeno-associated viral vectors in vivo. Mol Ther. 2013;21:796–805. doi: 10.1038/mt.2012.277. [DOI] [PMC free article] [PubMed] [Google Scholar]
  208. Tang J, Olive M, Pulmanausahakul R, et al. Human CD8+ cytotoxic T cell responses to adenovirus capsid proteins. Virology. 2006;350:312–322. doi: 10.1016/j.virol.2006.01.024. [DOI] [PubMed] [Google Scholar]
  209. Thaci B, Ulasov IV, Wainwright DA, Lesniak MS. The challenge for gene therapy: innate immune response to adenoviruses. Oncotarget. 2011;2:113–121. doi: 10.18632/oncotarget.231. [DOI] [PMC free article] [PubMed] [Google Scholar]
  210. Thwaite R, Pages G, Chillon M, Bosch A. AAVrh.10 immunogenicity in mice and humans. Relevance of antibody cross-reactivity in human gene therapy. Gene Ther. 2015;22:196–201. doi: 10.1038/gt.2014.103. [DOI] [PubMed] [Google Scholar]
  211. Toietta G, Mane VP, Norona WS et al (2005) Lifelong elimination of hyperbilirubinemia in the Gunn rat with a single injection of helper-dependent adenoviral vector. Proc Natl Acad Sci USA 102:3930–3935 [DOI] [PMC free article] [PubMed]
  212. Touchot N, Flume M. The payers’ perspective on gene therapies. Nat Biotechnol. 2015;33:902–904. doi: 10.1038/nbt.3332. [DOI] [PubMed] [Google Scholar]
  213. Trapani I, Colella P, Sommella A, et al. Effective delivery of large genes to the retina by dual AAV vectors. EMBO Mol Med. 2014;6:194–211. doi: 10.1002/emmm.201302948. [DOI] [PMC free article] [PubMed] [Google Scholar]
  214. Tse LV, Moller-Tank S, Asokan A. Strategies to circumvent humoral immunity to adeno-associated viral vectors. Expert Opin Biol Ther. 2015;15:845–855. doi: 10.1517/14712598.2015.1035645. [DOI] [PMC free article] [PubMed] [Google Scholar]
  215. Tseng YS, Agbandje-McKenna M. Mapping the AAV capsid host antibody response toward the development of second generation Gene delivery vectors. Front Immunol. 2014;5:9. doi: 10.3389/fimmu.2014.00009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  216. Unzu C, Sampedro A, Mauleon I, et al. Helper-dependent adenoviral liver gene therapy protects against induced attacks and corrects protein folding stress in acute intermittent porphyria mice. Hum Mol Genet. 2013;22:2929–2940. doi: 10.1093/hmg/ddt148. [DOI] [PubMed] [Google Scholar]
  217. Vercauteren K, Hoffman BE, Zolotukhin I, et al. Superior in vivo transduction of human hepatocytes using engineered AAV3 capsid. Mol Ther. 2016;24:1042–1049. doi: 10.1038/mt.2016.61. [DOI] [PMC free article] [PubMed] [Google Scholar]
  218. Verma IM, Somia N. Gene therapy - promises, problems and prospects. Nature. 1997;389:239–242. doi: 10.1038/38410. [DOI] [PubMed] [Google Scholar]
  219. Viecelli HM, Harbottle RP, Wong SP, et al. Treatment of phenylketonuria using minicircle-based naked-DNA gene transfer to murine liver. Hepatology. 2014;60:1035–1043. doi: 10.1002/hep.27104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  220. Vigant F, Descamps D, Jullienne B, et al. Substitution of hexon hypervariable region 5 of adenovirus serotype 5 abrogates blood factor binding and limits gene transfer to liver. Mol Ther. 2008;16:1474–1480. doi: 10.1038/mt.2008.132. [DOI] [PubMed] [Google Scholar]
  221. Waddington SN, McVey JH, Bhella D, et al. Adenovirus serotype 5 hexon mediates liver gene transfer. Cell. 2008;132:397–409. doi: 10.1016/j.cell.2008.01.016. [DOI] [PubMed] [Google Scholar]
  222. Wang L, Bell P, Somanathan S, et al. Comparative study of liver Gene transfer with AAV vectors based on natural and engineered AAV capsids. Mol Ther. 2015;23:1877–1887. doi: 10.1038/mt.2015.179. [DOI] [PMC free article] [PubMed] [Google Scholar]
  223. White GI, Monahan PE (2005) Gene therapy for hemophilia a. In: Lee CR, Berntrop E, Hoots K (eds) Textbook of hemophilia. Blackwell, Oxford, pp 226–228
  224. Wilson JM. Lessons learned from the gene therapy trial for ornithine transcarbamylase deficiency. Mol Genet Metab. 2009;96:151–157. doi: 10.1016/j.ymgme.2008.12.016. [DOI] [PubMed] [Google Scholar]
  225. Wilson EM, Bial J, Tarlow B, et al. Extensive double humanization of both liver and hematopoiesis in FRGN mice. Stem Cell Res. 2014;13:404–412. doi: 10.1016/j.scr.2014.08.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  226. Wirth T, Parker N, Yla-Herttuala S. History of gene therapy. Gene. 2013;525:162–169. doi: 10.1016/j.gene.2013.03.137. [DOI] [PubMed] [Google Scholar]
  227. Wold WS, Toth K. Adenovirus vectors for gene therapy, vaccination and cancer gene therapy. Curr Gene Ther. 2013;13:421–433. doi: 10.2174/1566523213666131125095046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  228. Wolff JA, Budker V. The mechanism of naked DNA uptake and expression. Adv Genet. 2005;54:3–20. doi: 10.1016/S0065-2660(05)54001-X. [DOI] [PubMed] [Google Scholar]
  229. Wu X, Liu G, Mu M et al (2016) Augmenter of liver regeneration gene therapy using a novel minicircle DNA vector alleviates liver fibrosis in rats. Hum Gene Ther 27:880–891 [DOI] [PubMed]
  230. Yan Z, Sun X, Feng Z et al (2015) Optimization of recombinant adeno-associated virus-mediated expression for large transgenes, using a synthetic promoter and tandem array enhancers. Hum Gene Ther 26:334–346 [DOI] [PMC free article] [PubMed]
  231. Yanez-Munoz RJ, Balaggan KS, MacNeil A, et al. Effective gene therapy with nonintegrating lentiviral vectors. Nat Med. 2006;12:348–353. doi: 10.1038/nm1365. [DOI] [PubMed] [Google Scholar]
  232. Yang Y, Ertl HC, Wilson JM. MHC class I-restricted cytotoxic T lymphocytes to viral antigens destroy hepatocytes in mice infected with E1-deleted recombinant adenoviruses. Immunity. 1994;1:433–442. doi: 10.1016/1074-7613(94)90074-4. [DOI] [PubMed] [Google Scholar]
  233. Yang Y, Wang L, Bell P, et al. A dual AAV system enables the Cas9-mediated correction of a metabolic liver disease in newborn mice. Nat Biotechnol. 2016;34:334–338. doi: 10.1038/nbt.3469. [DOI] [PMC free article] [PubMed] [Google Scholar]
  234. Yokoo T, Kamimura K, Abe H, et al. Liver-targeted hydrodynamic gene therapy: recent advances in the technique. World J Gastroenterol. 2016;22:8862–8868. doi: 10.3748/wjg.v22.i40.8862. [DOI] [PMC free article] [PubMed] [Google Scholar]
  235. Zaiss AK, Muruve DA. Immunity to adeno-associated virus vectors in animals and humans: a continued challenge. Gene Ther. 2008;15:808–816. doi: 10.1038/gt.2008.54. [DOI] [PubMed] [Google Scholar]
  236. Zhang HG, Zhou T, Yang P, Edwards CK, 3rd, Curiel DT, Mountz JD. Inhibition of tumor necrosis factor alpha decreases inflammation and prolongs adenovirus gene expression in lung and liver. Hum Gene Ther. 1998;9:1875–1884. doi: 10.1089/hum.1998.9.13-1875. [DOI] [PubMed] [Google Scholar]
  237. Zhu J, Huang X, Yang Y. Innate immune response to adenoviral vectors is mediated by both Toll-like receptor-dependent and -independent pathways. J Virol. 2007;81:3170–3180. doi: 10.1128/JVI.02192-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  238. Zufferey R, Dull T, Mandel RJ, et al. Self-inactivating lentivirus vector for safe and efficient in vivo gene delivery. J Virol. 1998;72:9873–9880. doi: 10.1128/jvi.72.12.9873-9880.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Journal of Inherited Metabolic Disease are provided here courtesy of Springer

RESOURCES