Abstract
Gene transfer has long been a compelling yet elusive therapeutic modality. First mainly considered for rare inherited disorders, gene therapy may open treatment opportunities for more challenging and complex diseases such as Alzheimer’s or Parkinson’s disease. Today, examples of striking clinical proof of concept, the first gene therapy drugs coming onto the market, and the emergence of powerful new molecular tools have broadened the number of avenues to target neurological disorders but have also highlighted safety concerns and technology gaps. The vector of choice for many nervous system targets currently is the adeno-associated viral (AAV) vector due to its desirable safety profile and strong neuronal tropism. In aggregate, the clinical success, the preclinical potential, and the technological innovation have made therapeutic AAV drug development a reality, particularly for nervous system disorders. Here, we discuss the rationale, opportunities, limitations, and progress in clinical AAV gene therapy.
Introduction
The concept of gene therapy was—to our knowledge-first explicitly articulated in 1947 (Keeler, 1947). What compelled the author Clyde Keeler mostly at the time were two aspects that differentiated genetic therapeutic paradigm from other treatment strategies: the ability to address the etiology of disease and the potential to induce a ‘‘permanent correction.” To date, these arguments remain the primary justification for pursuing gene therapy. Neurological diseases, even more than other classes of disorders, can benefit greatly from the unique advantages that gene therapy aims to provide. Targeting the root cause of disease is obvious for single-gene disorders but extends to common complex disease through a variety of genetic interventions. The permanency, or at least long-lived therapeutic effect, of a gene therapy is particularly attractive for organ systems such as peripheral and sensory organs, spinal cord, or brain, which are hard to access due to the inability of most drugs to cross physiological barriers such as the blood-brain barrier (BBB). Indeed, routes of administration (e.g., injections into the brain, cerebrospinal fluid [CSF], retina, or cochlea) that are clinically prohibitive for frequent injections of a traditional drug product with short acting half-life can be now considered more easily if only a single, long-lasting intervention is required.
In the 1970s, efforts in the field to develop genetic forms of medicine became more articulated; however, we had to wait for success in clinical gene therapy for many decades to come, particularly for in vivo gene therapy (Friedmann, 1992). By definition, in vivo gene therapy relies on the direct administration of the genetic drug to the patient as opposed to ex vivo approaches of therapeutic gene transfer that provide a more controlled laboratory environment for the gene transfer step. The progress in gene therapy was in lockstep with the increasing sophistication in recombinant techniques, human disease genetics, disease modeling, and molecular virology. These have now culminated in several powerful examples of clinical proof of concept such as inherited retinal disorders (IRDs) (Maguire et al., 2008) and spinal muscular atrophy (SMA) (Mendell et al., 2017).
The field of ex vivo gene therapy, in which tissues or cells are genetically modified in a clean room setting before they are administered to the patient, has advanced somewhat faster due to lessened safety concerns, at least in an autologous setting. In neurological disease, ex vivo gene therapy has previously led to demonstrable clinical benefit, e.g., in hematopoietic stem cell modification in several forms of leukodystrophy (Biffi et al., 2013; Cartier et al., 2009; Eichler et al., 2017). However, the generalizability of these approaches is limited, particularly for disease states due to cell-autonomous defects, by our current inability to transplant and functionally graft many cell types and tissues of the nervous system. The present review focuses on the clinical progress and development of in vivo therapeutic gene transfer.
Delivery is the key property that makes a therapy a gene therapy. The successful development of a drug relies on (1) a target, (2) a disease-modifying intervention, and (3) a delivery mechanism of said intervention to said target. While various types of interventions can be gene targeted, it is the delivery component that defines it as a gene therapy namely in that the drug substance itself is wholly or partly composed of genetic material. Formulating DNA or RNA to accomplish functional delivery to a cell, to a tissue, or into a complex physiological environment such as the CSF or bloodstream almost always requires a vector system. Indeed, unprotected (or naked) DNA or RNA injections have generally been limited in efficiency or actively rejected by the host. DNA and RNA degrading enzymatic activity, intrinsic stability, the inability to traverse lipid bilayers necessary to enter the cell and/or cellular nucleus, and the sensing mechanisms of the innate immune system are only some of the hostile elements that make in vivo gene delivery so challenging. While chemical modification, particularly of smaller polynucleotide chains, can provide some relief or protection to these attacks (Yin et al., 2014), formulation into a vector is most often needed to efficiently target a tissue at meaningful levels. Historically, non-viral and viral systems have been, and continue to be, explored. While each of the systems considered have unique benefits, for in vivo approaches that aim to provide long-lived to permanent therapeutic relief, adeno-associated virus (AAV) has emerged as a platform of increasing validation (Hastie and Samulski, 2015). Here, we discuss in detail the growing field of AAV gene transfer for neurological disorders.
AAV: Virus and Vector
AAV gene transfer is leveraging the properties of the AAV in the context of a replication-defective vector. As a wild-type virus, it is a member of the Parvoviridae family, characterized by its non-enveloped icosahedral capsid that carries a 4.7 kb singlestranded (ss) genome (Figure 1). AAVs were first identified as contaminants of laboratory adenovirus preparations. AAVs require ‘‘help’’ to complete a replication cycle, which can be provided by a co-infection with adenovirus, as well as a herpes virus and cellular stress. While adenoviruses can play a role in AAV’s life cycle, they are structurally and functionally highly distinct, both as a virus and in their utility as a gene transfer vector. The genome of AAV is flanked by inverted terminal repeats (ITRs), and three gene families are encoded: replication genes (Rep), structural capsid genes (Cap or VP1, −2, and −3), and a small viral co-factor required for assembly called assembly-activating protein (AAP) (Figure 1A) (Knipe and Howley, 2013).
Figure 1. AAVVirus and Vector.
(A) Genomic structure of the wild-type adeno associated virus (AAV), indicating the AAV2 genes encoding the replication (Rep), capsid (Cap), and assembly-associated protein (APP) products, flanked by inverted terminal repeat (ITR) secondary DNA hairpin structure.
(B) A recombinant AAV2-based vector is generated from the wild-type AAV2 genome by eliminating the open reading frames, solely retaining the flanking ITR sequences and replacing it with a transgene of interest upto a combined maximal sequence length of4.7 kb. These AAV2-based vector genomes (v.g.) are packaged in the protective capsid shell by coexpressing in trans the AAV gene products Rep, Cap, and AAP (in addition to essential gene products from helper virus such as human adenovirus type 5). Any AAV2-based vector genome can be packaged into various different types of AAV variant capsid (CapX) by cross-packaging in which the CapX gene is co-expressed with Rep from AAV2.
(C) The genomic structure of self-complementary (sc) AAV has within the same total packaging size of 4.7 kb two complementary copies of the same transgene linked by a third ITR sequence that was modified not to be resolved during viral amplification (black dot.
(D) The AAV virion is a multimeric protein composed of 60 subunits in a 20-facetted structure with an outer diameter of approximately 25 nm. All subunits are symmetrically integrated in the particle as dimers (red), trimers (blue), and pentamers (green) according to icosahedral symmetries.
AAVs are considered endemic in humans and have been isolated from several other species including the non-human primate (NHP), cow, goat, mouse, and even birds (Gao et al., 2005). The route of infection remains speculative; however, in humans and non-human primates, proviral AAV DNA was detected in liver and spleen, as well as a host of other tissues, suggestive of a stage of systemic viremia and indicative of its promiscuous tropism (Gao et al., 2003). In the absence of help, AAV genomes can reside in a latent state in the host cell. How this latency is maintained remains poorly understood; however, observationally, some of the genomes reside in an episomal state as concatamers (Duan et al., 1998). In careful in vitro studies, infection of a wild-type virus could lead to integration into the human 19q3 locus called AAVS1 (Kotin et al., 1990; Linden et al., 1996). This relative specificity is due to the sequence homology in the ITR region and integration in the host genome dependent on the presence of Rep gene products (Linden et al., 1996). Interestingly, a natural infection with AAV is not known to lead to disease, although some reports describe an association with hepatocellular carcinoma, which remains controversial (Nault et al., 2015, 2016; Srivastava and Carter, 2017).
Recombinant AAVs are generated by elimination of all open reading frames (ORFs) from the viral genome to render it into a replication-defective virus or vector (Figure 1B). The minimal coding that needs to be retained in cis to allow for an AAV-like genome to be packaged is restricted to the 145 bp within the ITRs that flank the transgene within AAV vectors (Hastie and Samulski, 2015). The absence of viral genes in the vector dramatically reduces safety concerns and frees up genetic cargo capacity for the insertion of a therapeutic transgene cassette. Indeed, the overall capacity of AAV to package an ITR-flanked genome productively is the approximate size of the wild-type AAV genome (i.e., 4.7 kb) (Figure 1B). Attempts to package larger genomes consistently lead to low titers and genome fragmentation (Dong et al., 2010; Lai et al., 2010; Wu et al., 2010). Due to its deficiency in autonomous replication, an AAV vector is produced in systems that complement the necessary components and cofactors for virion assembly, packaging, and amplification. Various systems have been developed to achieve this, particularly to support viral vector production at scale. The most common vector production is done through HEK293 transient transfection with multiple plasmids, namely one including the ITR-flanked cis transgene and others that encode the AAV rep and cap genes as well as a minimal set of adenoviral genes that provide help for the helper-dependent AAV to undergo viral replication (Figure 1B).
The desirable safety features of AAV as a virus and even more so as a replication-defective vector brought the system to the forefront. Decades in development, to date, AAV has been used in more than 200 human studies in which it was overall well tolerated in thousands of patients, further validating the favorable clinical profile of the platform (Ginn et al., 2018). The long development path of AAV, however, required technological advances to attain clinical success.
AAV Serotypes and the Race for Superior Capsids
Any gene delivery vector, when administered to the host, faces a hostile environment. The AAV capsid shell serves as a protection from nucleases and innate immune sensing inside and outside of the cell but also directs the tropism of the virion. Data indicate a determining role for the AAV capsid in its host response, cell attachment, cell entry, escape from the endosome, nuclear trafficking, release from the genome in the nucleus, and even its transcriptional state. Combined, the summation of these steps ultimately determines the safety and efficiency of an AAV administration following a particular route of administration (Pillay and Carette, 2017).
Variation in the primary sequence of the AAV capsid protein can alter its transduction efficiency dramatically. The first of those observations were made when exploring natural-occurring AAV serotypes. AAV1 to AAV6 were discovered in the 1960s and 1970s as contaminants of adenoviral preparations or viral cultures from tissue samples (Gao et al., 2005). These AAVs distinguished themselves serologically, hence the name serotype. AAV2, the first AAV to be molecularly cloned and rendered into a replication-defective vector, has been most extensively characterized for gene transfer applications. Other AAV serotypes were developed later as vectors, most often using a system called cross-packaging, which refers to the ability to package genomes flanked by AAV2 ITRs in any (non-AAV2) capsid by co-expressing it with the Rep genes from AAV2 during production (Figure 1B).
Early studies indicated that different AAV serotypes had a vastly distinct set of biological properties, including cell and tissue tropism (Auricchio et al., 2001). These observations sparked and continued to feed a buoying field of AAV capsid discovery and engineering. The goals of these efforts are to optimize AAV’s gene transfer functionality for particular uses (e.g., improve the ability to cross the BBB or minimize immunogenicity) or to introduce novel functions (e.g., allow transduction of a previously non-permissive cell type) (Grimm and Büning, 2017).
Methodologically, these discovery approaches can be broadly categorized in three different categories: (1) bio-mining, (2) rational design, and (3) bio-panning (also referred to as directed evolution). These methods have been reviewed extensively previously, however briefly: biomining seeks to rely on nature for novel capsid diversity, rational design aims to leverage structural and functional data on capsid domains to manipulate its utility, and in bio-panning, a laboratory-based library of AAV variants is subjected to a selective pressure to then be down-selected often in an iterative sequence (not unlike phage-display methods) (Deverman et al., 2018; Grimm and Büning, 2017; Van-denberghe et al., 2009). Importantly, these methods can be combined, e.g., in recent efforts to limit the complexity of a bio-panning library within particular functional domains, e.g., based on structural or evolutionary data (Deverman et al., 2018; Grimm and Büning, 2017; Grimm and Zolotukhin, 2015; Tse et al., 2017; Vandenberghe et al., 2009).
The Vector Genome as an Engine for Optimal Transgene Expression
With all attention on the viral capsid, the critical contribution of the vector genome cannot be forgotten. The minimal component that is required in an AAV vector genome are two ITRs that flank the transgene. Minimal ITRs have a docking site for the viral Rep protein that drive the vector genome into a preformed capsid particle. ITRs are not known to have strong intrinsic promoter activity; however, recent work has identified functional transcription factor binding sites that could act as enhancer elements, at least in liver (Logan et al., 2017). The genomic cargo of a vector transgene can be designed to carry any DNA, including a cDNA (gene augmentation or addition), a silencing RNA (gene silencing) (Borel et al., 2014), other RNA species (e.g., guide RNA, long non-coding RNA, or microRNA [miRNA]-binding sites) (Han et al., 2018; Karali et al., 2011; Platt et al., 2014), anti-sense oligonucleotides (Garanto et al., 2016), a DNA-or RNA-editing enzyme (gene editing) (Senís et al., 2014), a donor template for homologous repair (Wang et al., 2015), or docking sites for DNA-binding proteins such as transcription factors (Botta et al., 2016). When expression is desired, additional regulatory sequences, such as enhancer elements and promoters (including RNA Pol II and Pol III promoters), possible intronic sequence, and terminators such as polyadenylation signals, are required (Borel et al., 2014; Deverman et al., 2018). As noted before, all these elements should ideally not exceed 4.7 kb, the genome size of the wild-type AAV (Figure 1B).
For AAV, a ssDNA virus, to become transcriptionally active, its genome needs to undergo second-strand conversion, a key rate-limiting step in the transduction pathway (Fisher et al., 1996). This bottleneck was overcome by cleverly mutating one of the ITRs in which flanking genomes in the replication cycle are resolved into two complementary separate viral or vector genomes. The net effect of this manipulation is that a single capsid is able to package a single-stranded genome with two complementing transgene sequences, connected by a (mutated) ITR hinge. These double-stranded (ds) or self-complementary (sc) AAVs can only carry a little less than 50% of the normal AAV capacity (Figure 1C) (McCarty, 2008); however, in return, scAAVs have been shown in multiple tissue targets to lead to a faster onset of expression and overall significantly higher sustained levels of expression (McCarty, 2008). Particularly for hard-to-access tissues such as, for example, the central nervous system (CNS) by systemic injection, the advantage of scAAV can be critical in overcoming what appears a threshold-like effect for regular ssAAVs. It is generally thought that the success of the scAAV9 in clinical trials for SMA, as discussed later, is due to AAV9 enabling BBB crossing with scAAV technology providing it the local activity upon arrival (Mendell et al., 2017).
The Route of Administration: A Major Variable Determining the Safety and Efficacy
Gene transfer to the CNS or compartmentalized sensory organs, such as the eye and the ear, is challenging. Attaining an acceptable safety profile balanced with the appropriate level of cellular efficiency is highly dependent on the right combination of vector and route of delivery (Figure 2). The brain, eye, cochlea, and spinal cord, for example, are compartmentalized organs with physiological hurdles like the BBB greatly limiting access. Figure 2 provides an overview of the various routes of administration that are considered to target nervous system disorders in gene therapy. A subset of clinical studies using these various injection routes referred to in this section are summarized in Table 1.
Figure 2. In Vivo Routes of AAV Delivery to the Nervous System.
Local injections of vector to the eye or to the cochlea are preferably chosen for the treatment of neurosensory disorders, as a relatively small area needs to be targeted. Intraparenchymal infusion of AAV remains the most commonly used approach to deliver a therapeutic gene to the brain, even though other delivery strategies to the CSF (intra cerebroventricular, intracisternal, and intrathecal routes) or to the bloodstream could be advantageous for the treatment of multifocal diseases. Intranasal delivery is an alternative and non-invasive option potentially suitable to restore the levels of therapeutic lysosomal enzymes, which are secreted and can diffuse through the CNS. Finally, a few motor neuron diseases can be improved after intramuscular or intra-spinal AAV injection.
Table 1.
Ongoing Clinical Trials Using AAV for the Treatment of CNS Diseases
Current Clinical Programs of AAV Gene Therapy in the CNS (as of October 2018) | ||||||||
---|---|---|---|---|---|---|---|---|
Disease | Clinical Trial phase | Serotype | Transgene | Dose | Route of administration | ID ClinicalTrials.gov | Sponsor | References |
Lysosomal storage diseases | ||||||||
MPS I | Phase I (recruiting) | AAV6 | ZFN for safe insertion of hIDUA | 5 × 1012 vg/kg 1 × 1013 vg/kg 5 × 1013 vg/kg |
Intravenous | NCT02702115 | Sangamo Therapeutics | Harmatz et al., 2018 |
MPS II | Phase I (recruiting) | AAV6 | ZFN for safe insertion of hIDS | 5 × 1012 vg/kg 1 × 1013 vg/kg 5 × 1013 vg/kg |
Intravenous | NTC03041324 | Sangamo Therapeutics | Laoharawee et al., 2018 |
MPS IIIA | Phase I/II (recruiting) | AAV9 | hSGSH | 0.5 × 1013 vg/kg 1 × 1013 vg/kg 3 × 1013 vg/kg |
Intravenous | NCT02716246 | Abeona Therapeutics | N/A |
Phase II/III (recruiting) | AAVrh.10 | hSGSH | 7.2 × 1012vg total over 6 sites |
Intracerebral | NCT03612869 | LYSOGENE | Tardieu et al., 2014 | |
MPSIIIB | Phase I/II (active, not recruiting) | AAV5 | hNAGLU | 4 × 1012 vg total over 16 sites |
Intracerebral | NCT03300453 | UniQure Biopharma B.V. | Ellinwood et al., 2011 |
Phase I/II (recruiting) | AAV9 | hNAGLU | 2 × 1013 vg/kg 5 × 1013 vg/kg |
Intravenous | NCT03315182 | Nationwide Children’s Hospital | N/A | |
LINCL (Batten disease) | Phase I (active, not recruiting) | AAV2 | hCLN2 | 3 × 1012 vg total over 12 sites |
Intraparenchymal | NCT00151216 | Weill Medical College of Cornell University | Worgall et al., 2008 |
Phase I/II (active, not recruiting) | AAVrh.10 | hCLN2 | 9.0 × 1011 vg total 2.85 × 1011 vg total over 12 sites |
Intraparenchymal | NCT01161576 and NCT01414985 | Weill Medical College of Cornell University | N/A | |
Phase I/II (recruiting) | AAV9 | hCLN3 | Not disclosed (“low” and ‘‘high’’ doses) | Intrathecal | NCT03770572 | Nationwide Children’s Hospital | N/A | |
Phase I/II (recruiting) | AAV9 | hCLN6 | 1.5 × 1013 vg | Intrathecal | NCT02725580 | Nationwide Children’s Hospital | N/A | |
MLD | Phase I/II (active, not recruiting) | AAVrh.10 | hARSA | 1 × 1012 vg total 4 × 1012 vg total over 12 sites |
Intraparenchymal | NCT01801709 | Institut National de la Santé et de la Récherche Medicale, France | Colle et al., 2010; Zerah et al., 2015 |
Hereditary leukodystrophy | ||||||||
Canavan disease | Phase 1 recruiting | AAV2 | hASPA | 9 × 1011 vg over six sites | Intraparenchymal | NA | National Institute of Neurological Disorders and Stroke | Leone et al., 2012 |
Neurotransmitter disorder | ||||||||
AADC deficiency | Phase I/II (active, not recruiting) | AAV2 | hAADC | Not disclosed | Intraparenchymal (Putamen) | NCT01395641 | National Taiwan University Hospital | Hwu et al., 2012; Chien et al., 2017 |
Phase II (recruiting) - expansion of NCT01395641 | AAV2 | hAADC | 2.37 × 1011 vg total | Intraparenchymal (putamen) | NCT02926066 | National Taiwan University Hospital | ||
Phase I (recruiting) | AAV2 | hAADC | 1.3 × 1011 vg total | Intraparenchymal (SNc and VTA) | NCT02852213 | University of California, San Francisco | ||
Idiopathic age-related neurodegenerative diseases | ||||||||
Parkinson’s disease | Phase I/II (active, not recruiting) | AAV2 | NTN | 2.4 × 1012 vg total over 4 sites sham surgery |
Intraparenchymal (putamen and SNc) | NCT00985517 | Sangamo Therapeutics | Marks et al., 2010 |
Phase I (active, not recruiting) | AAV2 | hGDNF | 9 × 1010 vg total 3 × 1011 vg total 9 × 1011 vg total 3 × 1012 vg total |
Intraparenchymal + CED (putamen) | NCT01621581 | National Institute of Neurological Disorders and Stroke (NINDS) | Johnston et al., 2009; Su et al., 2009 | |
Phase I (active, not recruiting) | AAV2 | hAADC | 7.5 × 1011 vg total 1.5 × 1012 vg total 4.7 × 1012 vg total |
Intraparenchymal (striatum) | NCT01973543 | Voyager Therapeutics | Bankiewicz et al., 2006 | |
Phase II (recruiting) | AAV2 | hAADC | 2.5 × 1012 vg total versus sham |
Intraparenchymal (striatum) | NCT03562494 | Voyager Therapeutics | ||
Phase I (active, not recruiting) | AAV2 | hAADC | 9.4 × 1012 vg total | Intraparenchymal (striatum) | NCT03065192 | Voyager Therapeutics | ||
Phase I/II (recruiting) | AAV2 | hAADC | 3 × 1011 vg total 9 × 1011 vg total over 4 sites |
Intraparenchymal (putamen) | NCT02418598 | Jichi Medical University | Christine et al., 2009 | |
Alzheimer’s disease | Phase I | AAV2 | hNGF | 8.0 × 1010 vg total 2.5 × 1011 vg total 8.0 × 1011 vg total |
Intraparenchymal (basal forebrain region) | NCT00087789 | Ceregene | Rafii et al., 2014, 2018 |
Phase I (not yet recruiting) | AAVrh.10 | hAPOE2 | 8.0 × 1010 gc/kg 2.5 × 1011 gc/kg 8.0 × 10-gc/kg |
Intracisternal | NCT03634007 | Weill Medical College of Cornell University | Rosenberg et al., 2018 | |
Spinal cord disorders | ||||||||
SMA1 | Phase I (recruiting) | AAV9 | SMN | 6.0 × 1013 vg total 1.2 × 1014 vg total |
Intrathecal | NCT03381729 | AveXis | N/A |
Phase III (recruiting) | AAV9 | SMN | 1.1 × 1014 vg/kg | Intravenous | NCT03505099 | AveXis | Mendell et al., 2017 | |
Phase III (active, not recruiting) | AAV9 | SMN | Not disclosed | Intravenous | NCT03306277 | AveXis | ||
Phase III (recruiting) | AAV9 | SMN | Not disclosed | Intravenous | NCT03461289 | AveXis |
Active clinical interventional AAV studies listed on https://clinicaltrials.gov as of October 2018 are included while completed or terminated studies are not listed. Associated references correspond either to studies demonstrating the proof of concept of a strategy in animal models other than mice or to clinical results in patients already published.
Abbreviations: MPS, Mucopolysaccharidosis; LINCL, late infantile neuronal ceroidlipofuscinosis; MLD, metachromatic leukodystrophy; IDUA, alpha-L-iduronidase; IDS, iduronate2-sulfatase; SGSH, N-sulfoglucosamine sulfohydrolase; NAGLU, alpha-N-acetylglucosaminidase; ARSB, arylsulfatase B; CLN2, protein tripeptidyl peptidase-l; AADC, aromatic L-amino-acid decarboxylase; NTN, neurturin; GDNF, glial-cell-derived neurotrophic factor; NGF, nerve growth factor; APOE2, apolipoprotein 2; SMA1, spinal muscular atrophy type 1; SMN1, survival motor neuron 1; SNc, substantia nigra pars compacta; VTA, ventral tegmental area.
Local Routes of Administration: Intraparenchymal CNS, Ocular, and Cochlear
Local routes of vector administration have clear advantages over injections into the venous system or other fluid-filled compartments. It maximizes concentration and residence time of the gene transfer agent in close proximity to the target cells, avoids wide biodistribution, and thus limits the risk of immunogenicity or toxicities due to AAV components or ectopic expression of the transgene.
Neurosensory organs are particularly suited for local AAV delivery. Local injection to target ophthalmic disorders has progressed clinically the furthest, mainly driven by the fact that surgical access is relatively feasible, therapeutic interventions and their outcome can be monitored non-invasively, and the compartmentalized nature of the eye limits the dose and restricts systemic spread of the vector. Moreover, the retina and other targets in the eye have well-established mechanisms that promote immune privilege and thereby reduce the risk for inflammatory responses to the viral vector and transgene product. Various injection routes to administer to the eye are available and depend on the target cell type (Figure 2). Least invasive would be a topical administration on the cornea; however, AAV is only able to penetrate the tight corneal barriers inefficiently. Intracameral injections provide access to the anterior chamber and may be of interest to target the trabecular meshwork or the innervated cornea (Wang et al., 2017). Clinically, AAV programs have largely focused on degenerative retinal disorders, which are most efficiently targeted to the outer retina via subretinal injection and the inner retina and other cell types flanking the vitreous via intravitreal injection (Figure 2). In a subretinal injection, the vector is deposited between the outer retinal pigmented epithelial (RPE) cells and the neural retina. The injection of the vector bolus generates an iatrogenic retinal detachment referred to as a bleb, which resolves in animals and humans in a matter of hours. Via this route, most AAVs, within the bleb area, transduce the RPE and photoreceptor cells, a primary target for IRDs. Serotype difference influences the relative levels of transduction between RPEs and rod and cone photoreceptors, as well as limits transduction of other retinal target cells such as bipolar cells (Vandenberghe and Auricchio, 2012). AAV2 predominantly targets the RPE, a reason for its use, in Luxterna, the first FDA-approved AAV-based drug for an IRD with an enzyme dysfunction in the RPE (Russell et al., 2017). The efficiency of gene transfer following subretinal infusion is unparalleled; however, the injection procedure remains fairly complex and not routinely done clinically outside of cell and gene therapy. Some groups propose a suprachoroidal injection to provide access to the outer retina in an arguably less invasive manner (Peden et al., 2011) (Figure 2). The intravitreal injection is an attractive route that has clinically been pursued for targeting outer retinal cells, such as retinal ganglion cells, or the delivery of secreted therapeutic proteins (such as anti-VEGF agents) (Lukason et al., 2011; Trapani and Auricchio, 2018). The intravitreal injection procedure is clinically routine, is non-invasive, and has the potential to distribute across the whole retina rather than the restricted area within a subretinal bleb. However, particularly in large animal models and humans, AAV gene transfer following intravitreal injection is limited, requires high vector doses, and has a significantly increased risk of inflammation compared to subretinal injections (Miller and Vandenberghe, 2018). The appeal of intravitreal AAV gene transfer has led to a continued search for approaches and technologies that can increase its potency, reach both the inner and outer retina, and overcome the immunity obstacle to broaden its clinical use (Dalkara et al., 2013; Takahashi et al., 2017; Wassmer et al., 2017).
Like the eye, the ear presents an opportunity to deliver a vector bolus in the proximity of the neural target tissue in the cochlea at high concentration and minimal biodistribution. No clinical AAV programs have been pursued to date, in part due to the lack of precedent for intra-cochlear drug delivery; however, experimentally, Figure 2 describes the various routes that have been explored for AAV gene transfer preclinical studies (Holt and Vandenberghe, 2012; Tao et al., 2018; Yoshimura et al., 2018). In the only cochlear gene transfer study to date that has been performed in humans (NCT02132130), human adenoviral vector was injected via an oval window procedure.
Local delivery to the CNS involves a surgical procedure during which a subject is being anesthetized and restrained within a stereotaxic frame while a needle tip or a flexible fused silica catheter is directly inserted in the parenchyma through burr holes drilled in the skull (Figure 2). For direct AAV injection in small animal models (mice and rats), stereotaxic coordinates are generally calculated from a reference point on the skull (bregma) accordingly to a brain atlas, while real-time MRI guidance systems are used for increased accuracy in larger species such as NHPs and humans (Eberling et al., 2008; Fiandaca et al., 2009; Hadaczek et al., 2006). Most AAV serotypes generally show excellent neuronal tropism (Cearley and Wolfe, 2007), while AAV5 also efficiently targets astrocytes (Davidson et al., 2000). By contrast, transduction of microglial cells appears much more challenging, with only one report so far achieving this ‘‘tour de force’’ using a capsid-modified AAV6 (Rosario et al., 2016). AAV intraparenchymal injections have been successful in numerous pre-clinical studies using mouse models of lysosomal storage diseases (LSDs), Alzheimer’s (AD), Parkinson’s (PD), and Huntington’s (HD) diseases (detailed in two excellent reviews, Hocquemiller et al., 2016; Choudhury et al., 2017). In those small animal models, broad transduction is achieved with only a few microliters of vector locally injected (generally 1–2 × 1010 vg per site). Safe and widespread correction has also been reported in larger species, including feline models of GM1 and GM2 gangliosidoses (using AAV1 or AAVrh.8; McCurdy et al., 2015), dog models of Mucopolysaccharidoses types I and IIIB (using AAV5; Ellinwood et al., 2011), and NHPs after induction of a lesion in the putamen with MPTP (1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine) to mimic Parkinsonian neurodegeneration (with sustained clinical improvement observed over 8 years after delivery of AAV2-hAADC; Hadaczek et al., 2010). For other conditions for which no large animal models are available to reproduce the clinical symptoms of human diseases, safety and transduction efficacy studies after AAV intracerebral injection have been carried on in healthy NHPs and generally well tolerated (Colle et al., 2010; McBride et al., 2011; Zerah et al., 2015, Ciron et al., 2009; Sondhi et al., 2012). This modality of vector delivery has also been the first in humans for a CNS application (Janson et al., 2002) and remains the norm in most clinical programs so far (see Table 1). For motor neuron disorders, multilevel injections in the spinal cord parenchyma have been successful to deliver therapeutic agents in amyotrophic lateral sclerosis (ALS) mouse models (Azzouz et al., 2000; Franz et al., 2009; Hardcastle et al., 2018). Translation to larger mammals and humans has been challenging due to the risky surgical intervention and the focal nature of the resulting transgene expression.
Besides the risk of viral or bacterial infection, hemorrhage, and edema inherent to any neurosurgical intervention, the main limitation associated with AAV stereotaxic injection is the restricted spread of the vector in the parenchyma, especially for therapeutic applications to diseases affecting large areas of the CNS or peripheral nervous system (PNS). Longdistance anterograde and retrograde axonal transport of certain AAV serotypes (AAV1, −6, −8, and −9) has been reported and could promote dissemination of the viral particles across anatomically connected areas of the brain (Castle et al., 2014; Löw et al., 2013; San Sebastian et al., 2013; Zingg et al., 2017). Convection-enhanced delivery (CED), which consists of the application of a certain pressure during infusion to enhance the vector diffusion, is another option (Bankiewicz et al., 2000; Hadaczek et al., 2006). Other attempts to potentiate AAV transduction after local delivery in the brain include co-infusion of factors such as heparin (Mastakov et al., 2002) or mannitol (Mastakov et al., 2001), in possible combination with CED (Carty et al., 2010). Finally, in specific neurological contexts such as neuropathic LSDs, the release of therapeutic lysosomal enzymes from transduced neurons and their uptake by distal cells leads to an expansion of the corrective effect of the gene therapy to larger areas than where AAV particles diffuse. This cross-correction process depends upon the presence of mannose 6-phophate receptors on recipient cells and is conserved among mammals. Considering that relatively low residual enzymatic activity can be clinically relevant (Leinekugel et al., 1992), intracerebral delivery of therapeutic genes for the treatment of infantile LSDs remains the main application for AAV-based gene therapy in the clinic (Table 1).
Delivery to the CSF via Intracerebroventricular, Intracisternal Delivery, or Intrathecal Routes
Delivery to the CSF allows for more widespread gene transfer throughout the brain and spinal cord (Hardcastle et al., 2018) while limiting systemic biodistribution (Figure 2). However, the presence of tight junctions between ependymal cells restricts such application to certain AAV serotypes (such as AAV1, −2, −4, −5, −8, −9, rh10, and PHP.B) able to extravagate through this protective cell layer (Hinderer et al., 2018a). Importantly, the age of the animals at the time of intervention influences the transduction profile following AAV infusion in the CSF: for example, a comparative study demonstrated that intracerebroventricular (i.c.v.) injection of AAV8 and AAV9 resulted in widespread biodistribution in the brain as compared with other serotypes (AAV1, −2, −5, and −7) when administered on postnatal day 0 (P0) and that the transduction profile tended to shift from neurons to glial cells between P0 and P1-P2 (Chakrabarty et al., 2013).
In pre-clinical studies using rodents, i.c.v. or intracisternal (IC) infusions of AAV in neonates and adult mice lead to widespread expression of transgenes across the cerebral tissue and the spinal cord, a strategy that has been successful in correcting symptoms of MPS I (AAV8-hIDUA; Wolf et al., 2011), II (AAV9.hIDS; Hinderer et al., 2016), IIIA (AAV9.CAG; Haurigot et al., 2013), VII (AAV1-GUSB; Passini et al., 2003), GM1 gangliosidosis (Broekman et al., 2007), SMA (AAV9.SMN1; Armbruster et al., 2016), MDL (AAV1-hARSA; Hironaka et al., 2015), ALS (AAV.miRNA-SOD1; Dirren et al., 2015), and even AD (Levites et al., 2006). Similar strategy has been attempted in larger species such as dogs for the treatment of ceroid lipofuscinosis (Biferi et al., 2017) and MPS IIIA (Haurigot et al., 2013) and cat models of Sandhoff disease (Rockwell et al., 2015). Interestingly, transduction of the ependymal cell layer and/or the choroid plexus by AAV2 or AAV4 also improved disease symptoms without requiring the transport of viral particles across this biological barrier (Liu et al., 2005, Hudry et al., 2013, Benraiss et al., 2013). In those cases, it is assumed that the ependymal and choroid plexus cells act as therapeutic reservoirs for the secretion of soluble enzymes or neurotrophic factors throughout the cerebral tissue. Whether or not long-term gene expression can be achieved remains unknown, especially because of the relatively fast turnover rate of the ependymal cells (Chauhan and Lewis, 1979).
By comparison, intrathecal (IT) delivery of AAV vectors at the level of the lumbar cord is easily translatable to the clinic considering that this space is routinely accessed into via lumbar puncture. Additionally, the insertion of a catheter through the lower section of the spinal cord can eventually be advanced rostrally to the cisterna magna with limited risks. Such an approach has been used for the preclinical evaluation of therapeutic candidates (serine histogranin, endomorphin1) for the relief of neuropathic chronic pain symptoms (using an AAV8; Jergova et al., 2017) and to correct ALS neurodegenerative processes in SOD1 mice (using an AAVrh.10; Li et al., 2017). A single IT injection of AAV.hGAA also led to neurologic and cardiac improvements in Pompe disease (Hordeaux et al., 2017), while similar central and peripheral benefits could be demonstrated after infusion of AAV9, AAVrh.10, and AAV.Olig001 in young Krabbe disease mice (10–11 days old; Karumuthil-Melethil et al., 2016). Overall, the transduction efficacy appears to be rather equivalent in adult mice when comparing different routes of administration in the CSF (Bey et al., 2017), but infusion of AAV9 in the cisterna magna or in the lateral ventricle has been shown to be far more potent at transducing the brain and spinal cord of cynomolgus macaques or dogs as compared with IT injection by lumbar puncture (Hinderer et al., 2018a). Whether or not placing the subject in the Trendelenburg position right after infusion can potentiate the efficacy of brain transduction by this procedure remains debated (Meyer et al., 2015).
Intranasal Delivery for Brain Transduction
While the primary indication for intranasal AAV delivery has been to target the respiratory tract, this non-invasive gene transfer strategy has also been recently applied for the treatment of MPS I (Belur et al., 2017) (Figure 2). Adult IDUA-deficient mice were instilled intranasally with an AAV9.IDUA, and a complete restoration of enzyme activity to normal levels was noticed 5 months post treatment in all parts of the brain, even though the area of AAV transduction remained localized to the olfactory bulb. This approach may be suitable for other LSDs considering the possible trafficking of therapeutic lysosomal enzymes to the CNS via the olfactory and trigeminal tracks innervating the nasal passages (Wolf et al., 2012). Intranasal delivery of an AAV vector driving expression of the neurotrophic factor BDNF fused with cell-penetrating peptides (TAT and HA2) was also associated with anti-depressant effects in mice and is considered a promising therapy for major depressive disorder (Ma et al., 2016).
Intramuscular Injection and Nerve Delivery to the Spinal Cord
Targeting the spinal cord can be achieved via remote delivery in the muscle and relies on the ability of certain AAV serotypes to hijack the axonal transport machinery and travel along the nerve following a centripetal way (Boulis et al., 2003; Towne et al., 2010) (Figure 2). AAV1, AAV6, and, more recently, AAV9 have led to efficient targeting of motor neurons after intramuscular infusion in rodents (Benkhelifa-Ziyyat et al., 2013; Hollis Ii et al., 2008; Towne et al., 2010). Infusion of a ssAAV6 in the gastrocnemius muscle of African green monkeys has been attempted and transduced about 50% of motor neurons in the ventral horn of the ipsilateral-injected side of the spinal cord (1.3 × 1012 vg injected; Towne et al., 2010). AAV intramuscular injections appear especially suited for the treatment of motor neuron diseases such as ALS, SMA, or pain (Benkhelifa-Ziyyat et al., 2013; Wang et al., 2002). Unfortunately, scaling up the dose of vector to treat larger animals has proven challenging, the spread of the vector remaining limited and localized in the lumbar segment of the spinal cord. Additionally, early dysfunction of the axonal transport machinery in those diseases may prevent efficient AAV transduction (Bilsland et al., 2010; Fischer et al., 2004).
Intravenous Delivery for Non-invasive and Broad Transduction of the CNS
The BBB efficiently isolates the CNS from the peripheral compartment and is a complex biological barrier constituted by successive layers of brain endothelial cells coherently connected by tight junctions, pericytes, and astrocytic endfeet (Figure 2). Numerous efforts have been made to try to disrupt the BBB in order to allow the transfer of therapeutic agents. Infusion of mannitol, for example, has been used to potentiate the transduction of neural tissue after intravenous delivery of AAV (Mastakov et al., 2001; McCarty et al., 2009), an approach somewhat successful but associated with significant adverse effects.
Because of those challenges, the characterization by Foust and colleagues of an AAV (AAV9) able to naturally bypass this “impermeable” biological barrier constitutes a definite breakthrough in the field (Foust et al., 2009). For the first time, large-scale transduction of the entire brain and widespread expression of therapeutic genes in the CNS became achievable after a single intravenous injection of vector, a feature especially advantageous for the treatment of multifocal diseases (Foust et al., 2010; Mattar et al., 2013; Murrey et al., 2014). Indeed, illustrated in Figure 3, the combination of scAAV and AAV9 at high dose can result in widespread transduction of spinal cord and the CNS following an IV injection (Hudry et al., 2018). This finding served as the basis of many preclinical and clinical programs described in more detail below. It also spurred further technology exploration. Indeed, other AAV serotypes appear to have qualitatively similar transduction abilities (Yang et al., 2014; Zhang et al., 2011), even though those remain less well characterized across various disease and animal models than AAV9. Capsid engineering studies have further led to AAV variants that quantitatively appear superior to AAV9 such as AAV.PHP.B and AAV-B1 (Choudhury et al., 2016a; Deverman et al., 2016) and AAV-AS (Choudhury et al., 2016b), at least in the animal model tested in the studies (AAV.PHP.B appears to not sustain that superior ability in other species; Challis et al., 2019; Hordeaux et al., 2018). Of note, most of the successful attempts of systemic AAV9 delivery have been done after packaging self-complementary genomes (Figure 1C). Anc80L65, a fully rationally designed AAV, has been shown to—at least in part—overcome the need for a self-complementary genome to accomplish substantial CNS transduction following systemic injection (Deverman et al., 2016; Hudry et al., 2018; Zinn et al., 2015).
Figure 3. AAV9 Brain Transduction after Intravenous Delivery in C57BL/6.
Representative images of eGFP fluorescence signal (and DAPI) across three different coronal sections across the brain after intravenous injection of AAV9 harboring a self-complementary genome (4 × 1013 vg/kg) expressing CMV.eGFP (Hudry et al., 2018) at the level of the striatum (upper panel), hippocampus (middle panel), or cerebellum (bottom panel). C57BL/6 animals were injected via tail vein at 6–8 weeks of age. Scale bar, 1,000 µm.
Due to its extensive validation studies across models and by multiple groups, AAV9 remains the gold standard serotype for CNS targeting after systemic delivery and has been extensively used with minimal signs of peripheral or central toxicity (Murrey et al., 2014; Pleger et al., 2011; Saraiva et al., 2016). Interestingly, notwithstanding some discrepancies between the reported studies, the transduction profile in the brain in animals has been reported to shift accordingly to the age of treatment: while neurons are preferentially targeted after delivery in neonates, astroglia mostly express the transgene after infusion in adult mice (Deverman et al., 2016; Foust et al., 2009; Hudry et al., 2016). Systemic injection of AAV9 in newborn or adult mice transduced motor neurons (MNs) in the spinal cord, up to 78% in lower spinal MNs according to Wang and colleagues in newborn mice (Wang et al., 2010) but with declining efficiencies later in life (Duque et al., 2009; Gray et al., 2011). AAV9 studies in NHPs confirmed those initial findings, showing efficient transduction of neurons after intravenous infusion in neonatal rhesus macaques, while glial targeting was prominent after administration in juvenile subjects (Gray et al., 2011; Mattar et al., 2013). MNs were consistently targeted regardless of the age of the subject. Preclinical results have been rather positive so far (for GM1 gangliosidosis [Weismann et al., 2015], Sandhoff disease [Walia et al., 2015], MPS IIIB [Fu et al., 2011; Murrey et al., 2014], AD [Iwata et al., 2013], HD [Dufour et al., 2014], SMA [Benkhelifa-Ziyyat et al., 2013], or ALS [Yamashita et al., 2013]). Importantly, the first clinical report of gene-replacement therapy for SMA type I has recently been published and showed remarkable safety and efficacy of this approach (Mendell et al., 2017).
Even though AAV peripheral delivery is technically easy and non-invasive, important challenges complicating clinical application remain: (1) large doses of vector are required to achieve relevant transduction in the CNS and improve disease symptoms; (2) intravenous administration exposes the vector to anti-AAV antibodies in those individuals that were previously exposed to AAV; (3) as most AAVs are hepatotropic, most of the vector will distribute to the liver, increasing the risk for hepatotoxicity; and (4) ectopic expression of the transgene in peripheral tissues may further pose safety risks.
In summary, the choice of the route of administration is a critical one yet requires a balancing of multiple parameters along a risk-benefit equation. Historically, local routes of injection have been preferred as they minimize the safety considerations while maximizing delivery. Driven by the desire to overcome the invasiveness of the local administration procedure and/or the clinical need to target more broadly than what a local injection route can deliver, novel technologies were, and often still are, needed that enable meaningful levels of gene transfer via these routes. These approaches, at the current stage, almost always require higher levels of dosing, increasing the safety risk and putting significant burden on AAV manufacturing. In certain cases, various routes of administration are combined when either technologies are too limiting or the disease presentation (e.g., in syndromic cases) requires multiple organs to be targeted (Biferi et al., 2017; Gurda et al., 2016).
Safety Consideration in AAV Gene Therapy
A primary driver for AAV’s adoption in gene therapy approaches has been its remarkable safety record in preclinical and clinical studies. Even at high doses, traditional toxicology assessment is relatively uneventful. However, one cannot be complacent as the first reports of dose-limiting toxicities using AAV emerge. Indeed, in a recent study in piglets and NHPs, an AAV variant administered i.v. at high dose led to severe toxicity that resulted in death or necessitated euthanasia days after vector administration. Observations included elevated transaminases, degeneration of dorsal root ganglia, impaired ambulation, proprioceptive deficits, and ataxia (Hinderer et al., 2018b). AAV gene therapies are complex biological agents for which the clinical experience of the platform remains in its infancy. Moreover, host responses can be unpredictable and dependent on disease state, route and compartment of administration, genetic predispositions, or other idiosyncratic variables.
Host Response
Immunological responses to a gene therapy can significantly impact the safety and longevity of any gene therapy. Possible sources of antigenicity are the vector components, the expressed transgene, and contaminants of the vector preparation. The immunogenicity can be innate and/or adaptive in nature.
Gene therapies, like with other biologics, can be complicated by pre-existing immunity (PEI). PEI is due to prior exposure of the host to one of components in the gene therapy and the lasting memory immune responses to which the initial exposure may lead. Since AAV is endemic in humans, PEI complicates the development of gene therapies at various levels as most elements that make up the gene therapy are naturally derived and thereby possible to have pre-exposed the recipient. Individuals with memory B cell responses have circulating levels of anti-AAV neutralizing antibodies (NABs) that can, even at low levels, prevent the AAV particle from reaching its intended tissue or cell target. The most immediate impact of AAV NABs is therefore on (the lack of) efficacy, not safety. Sequential AAV injections may suffer from NABs elicited by a prior administration, and re-administration is therefore currently not considered a feasible option (Amado et al., 2010). PEI to AAV can also manifest itself through a memory T cell response; indeed, the reactivation of memory T cell responses has been hypothesized to play a role in hepatotoxicity and the loss of transgene expression that has been observed in several AAV clinical trials. The prevalence and impact of PEI to AAV are significant but difficult to gauge quantitatively due to differences in the detection assays and the vector serotypes tested between populations (e.g., age, clinical history, geographical location). General concepts on immunological and other safety considerations for AAVs have been reviewed extensively elsewhere (Mingozzi and High, 2013; Van-damme et al., 2017; Vandenberghe and Wilson, 2007).
PEI is not solely an issue of vector capsid, as it can also be directed to the expressed transgene. An example of this in a traditional gene augmentation context is LSD for which enzyme replacement therapies are available (Wang et al., 2008). During these treatments, some patients develop antibody responses to the injected enzyme. If these patients later are receiving a gene therapy expressing the same enzyme, these NABs constitute (drug-induced) PEI in the context of the gene therapy. Another example relates to the use of CRISPR genome editing, as the bacterial CRISPR enzyme itself is an antigen that humans are exposed to throughout life (via a common bacterial infections) (Charlesworth et al., 2018).
Fortunately, AAV gene therapies for neurological disease have been somewhat spared from major inflammatory adversity, likely due to the immune privilege that many of the compartments in the nervous system benefit from (Carson et al., 2006; Forrester et al., 2018). Brain, spinal cord, eye, and cochlea have limited access to circulating antibodies or infiltrating immune cells and, in certain cases, mechanisms to actively dampen immune response (Forrester et al., 2018; Stein-Streilein, 2013). NABs due to PEI, for example, are significantly reduced in the ocular vitreous or CSF. Intraparenchymal CNS or subretinal injections even have less exposure to NABs and likely are not affected, even if the host is highly serologically positive or has previously received an AAV injection (e.g., when administering a retinal gene therapy in both eyes) (Amado et al., 2010; Gray et al., 2013). Indeed, clinical programs that relied on direct intraparenchymal or intra-CSF injections overall confirmed the safety and tolerability of these approaches (detailed in Table 1 and in the section ‘‘Clinical Progress: Illustrative Case Studies’’), with little signs of inflammation reported and only a small subset of patients eventually developing a mild immune reaction against the vector capsid. Similarly, in the eye, programs using a subretinal route of injection have overall safely progressed in the clinical studies, presumably due to the local immune privilege of the retina, the small dose, and the containment of the vector in the subretinal bleb (Trapani and Auricchio, 2018). Nonetheless, immune privilege is a relative concept, as in higher animal studies inflammatory responses have been noted particularly when higher doses are administered, possibly due to innate immune sensing in the retina or leakage of virus into the vitreous due to the subretinal injection procedure (Khabou et al., 2018; Reichel et al., 2017; Vandenberghe et al., 2011). Direct intravitreal, which conceptually has noted advantages over subretinal injection, illustrates the limits of immune privilege even more. Contrary to many other commonly used protein-based drugs, AAV injected in the vitreous was found to be inflammatory, even at moderate doses (Miller and Vandenberghe, 2018). These responses remain poorly understood, mainly because they are only evident in larger animal models and human studies (Cukras et al., 2018; Maclachlan et al., 2011), and while responsive to pharmacological immune suppression, further research is needed to adequately avoid or mitigate these responses.
Systemic AAV administration with the goal to broadly transduce the neural tissue is evidently more challenging than local injection routes. Most studies are forced to exclude patients with detectable levels of NABs to AAV, as it is expected they will not benefit from the gene therapy. Even then, AAV clinical studies for hemophilia showed that seronegative subjects developed transaminase elevations weeks after AAV administration, which in some patients was concomitant with a rise in AAV capsid-specific CD8+ cells and the loss of transgene expression (Manno et al., 2006; Mingozzi et al., 2007; Nathwani et al., 2011). In the SMA study in which AAV9 was given IV at high dose, asymptomatic elevation of transaminases was observed in some of the newborn subjects, who fortunately responded to prednisolone treatment (Mendell et al., 2017).
Cellular immune response toward heterologous or even autologous gene products poses another concern to take into account when considering AAV gene transfer. Even after direct intraparenchymal or intra-CSF injection, a few pre-clinical studies in dog models of MPS IIIA disease, Sanfilippo, and Hurler syndromes have actually showed that a robust inflammatory response can arise from AAV-driven expression of human transgenes, while no such response was triggered when the canine cDNAs were cloned in the vector backbone (Ellinwood et al., 2011; Haurigot et al., 2013).
Ectopic expression of the transgene product can be a source of antigenicity or toxicity, an issue that warrants the development of strategies to restrict expression of the therapeutic protein to targeted tissues. Capsid modifications can be made to reduce liver targeting and promote transduction of the neural tissue (Castle et al., 2016; Kanaan et al., 2017). Another successful strategy has been to tailor the AAV genome and insert cell-specific promoters allowing specific expression in defined cell populations such as neurons (synapsin or CamKII promoters; Jackson et al., 2016; McLean et al., 2014; Watakabe et al., 2015), astrocytes (GFAP or ALDH1/1 promoters; Dashkoff et al., 2016; Koh et al., 2017), or oligodendrocytes (using a myelin basic protein promoter; Georgiou et al., 2017; Kennedy and Rinholm, 2017). miRNA-specific target sequences have also been cloned in the viral genome (3’ UTR) to abolish off-target expression in undesirable cell types (Taschenberger et al., 2017).
Currently utilized clinical strategies to side step or overcome immune concerns brought on by AAV gene therapies include the exclusion of patients with elevated anti-AAV antibodies from enrolling in clinical trials, administering in compartments that are thought to be immune privileged, including short-term immunosuppression regimens (e.g., oral prednisone; Mendell et al., 2017; Nathwani et al., 2011), or limiting the vector dose (Mendell et al., 2017). While these strategies have overall proven to be adequate in the ongoing clinical studies, it is clear they do not provide a solution for many of the outstanding concerns in the field. How can we effectively administer an AAV to a seropositive patient? Can we do repeat administrations of AAV over time? How do we safely perform AAV gene transfer with an antigen that is foreign to patient, such as when a patient has null mutations in an autosomal recessive disease gene? These and other questions remain the topic of investigation and technology development in the laboratory (reviewed in Mingozzi and High, 2013).
Genotoxicity and Insertional Mutagenesis
Contrary to retroviral vectors, recombinant AAV is not known to actively integrate into the host genome but rather resides in a non-replicating episomal state upon transduction in the target cell (Schnepp et al., 2005). Indeed, since no mechanism actively duplicates the transgene during mitosis, its expression is eventually lost in a dividing cell population. However, infrequent integration of the AAV transgene into the host genome can occur, with a bias toward active genomic regions (Nakai et al., 2003). The risk for generating insertional mutagenesis therefore does exist and was recently reviewed in detail (Chandler et al., 2017). While this safety issue has not yet been reported in human patients, increased incidence of hepatocellular carcinoma (HCC) has been observed after systemic delivery of AAV2 and AAV9 in newborn MPSVII mice (AAV2) and Sandhoff disease model as the animals get older (between 43 and 52 weeks post injection) (Donsante et al., 2007; Walia et al., 2015; Zhong et al., 2013). The incidence of those events reaches 30%−60% after injection of AAV2 at a dose of 1.5 × 1011 vg and 80% with AAV9 at the much higher dose of 2.5 × 1014 vg with similar observations in non-transgenic littermates. Further investigation of the genetic material of the tumors revealed the presence of recombinant AAV genome integrated in a specific locus of the murine chromosome 12 (Rian locus), leading to dysregulation of expression of the surrounding genes (Donsante et al., 2007; Wang et al., 2012). While those findings have recently been reproduced in a large study using various different serotypes and longer time points (up to 24 months) (Chandler et al., 2015), two additional reports eventually could not reproduce those findings (Bell et al., 2005; Li et al., 2011), but the developmental timing of injections was different, suggesting age-difference susceptibility to rAAV genotoxicity. Additionally, a detailed genomic analysis demonstrated that the vector integration sites were mapped in Mir341 (within the Rian locus), which only exists in rodents (Chandler et al., 2015). In a select number of subjects with HCC that were not AAV gene therapy recipients, genomic analysis of HCC biopsies identified AAV2 genomes to be integrated in oncogenic driver genes (Nault et al., 2015). Whether or not the reported pathogenicity of wild-type AAV2 is relevant to recombinant AAV use in the clinics is unclear (Gil-Farina et al., 2016). In the specific case of gene transfer applied to neurological disorders, insertional mutagenesis may only be a concern when high doses of vector are injected systemically. It also implies that local injections of small amounts of AAV in the parenchyma or CSF will be safe, even though one paper reported a possible integration of AAV in non-dividing cells such as neurons (Wu et al., 1998).
Clinical Progress: Illustrative Case Studies
Clinical gene therapy has come a long way to overcome perceived and real concerns about safety and its potential to markedly intervene in disease in humans, and today the demonstrated proof of concept of the clinical utility of gene therapy is clear. Remarkably, two of those pioneering programs are in nervous system disorders and both lean on AAV. The FDA in 2017 and the European Medicines Agency (EMA) in 2018 approved for the market Luxturna (voretigene neparvovec-rzyl), a subretinally injected AAV encoding the cDNA for RPE65 for treatment of patients with confirmed biallelic RPE65 mutation-associated retinal dystrophy that leads to vision loss and may cause complete blindness in certain patients (Russell et al., 2017). In late 2018, an application was submitted to the FDA, the EMA, as well as Japanese regulators for the market approval of AVXS-101 or Zolgensma (onasemnogene abeparvovec), the Novartis/AveXis AAV gene therapy for SMA (Mendell et al., 2017). If approved—at the earliest in the first half of 2019—it would become the second AAV product currently on the market.
Here, we present several case studies of progress toward or in the clinic of AAV gene therapies targeting the nervous system, with a specific emphasis on diseases of the CNS (summarized in Table 1). For an in-depth overview of the progress and clinical studies in the field of ophthalmology, specifically for inherited retinal disorders, we refer to a number of recent reviews on the subject (DiCarlo et al., 2018; Trapani and Auricchio, 2018). Only a few clinical studies have explored AAV gene transfer applied to inherited peripheral neuropathies, with only one phase I/IIa clinical trial registered so far for the treatment of Charcot-Marie-Tooth disease (NCT03520751) and of giant axonal neuropathy (that affects both PNS and CNS, NCT02362438), yet without any reported results.
Neurosensory Disorders
In some ways, it is remarkable that the first convincing demonstration of the clinical impact of in vivo gene therapy was one in a neurosensory disorder; in 2008, the now FDA-approved Luxterna was shown to benefit patients suffering from a type of Leber congenital amaurosis due to mutations in the gene RPE65 (Maguire et al., 2008). This was remarkable, because, historically, gene therapy was primarily considered for disorders with significant mortality, such as muscular dystrophy or genetic forms of immunodeficiency (Miller, 1992). In hindsight, the rationale for the pursuit for therapeutic gene transfer in neurosensory disorders appears obvious: the tissue target is compartmentalized, warranting a local injection that can be positioned in close proximity to the target cells and thereby limits systemic exposure, a smaller dose requirement, and lessened safety concerns. Approaches in the retina have been on the forefront for the past two decades, resulting in Luxterna and multiple other gene therapy programs for blinding disorders currently in the clinic (reviewed by DiCarlo et al., 2018; Trapani and Auricchio, 2018). The majority of the clinical programs have been based on AAV due to its high gene transfer efficiency to retinal pigment epithelial cells and photoreceptors following a subretinal injection and the availability of many AAV variants that demonstrate altered specificity for the various retinal cell types (Vandenberghe and Auricchio, 2012). The success and growing level of comfort around the use of AAV in the retina continues to push boundaries with the first in vivo experimental clinical trials using optogenetics and gene-editing techniques, such as CRISPR, to be targeting the retina and building on the safety and potency of AAV (DiCarlo et al., 2018).
An emerging field that draws on analogies with retinal gene therapy is focused on therapeutic gene transfer for hearing and balance disorders (Géléoc and Holt, 2014). Indeed, the cochlea is also a well-confined space allowing for localized infusions at low doses. Several important distinctions, however, may explain the fact that hearing gene therapy has not seen to date the clinical advances seen in ophthalmology: only recently have novel AAV technologies emerged that improve targeting efficiencies (György et al., 2017, 2018; Landegger et al., 2017); drug delivery into the inner ear is not routine; hearing aids and cochlear implants are a standard of care that can provide therapeutic relief to patients; and lastly, many inherited, particularly autosomal recessive forms, hearing loss disorders likely have either no or only a very short therapeutic window of intervention, thus complicating the translation of gene therapies for these disorders. Nonetheless, the large remaining unmet need and the increasing comfort around AAV gene therapy development is making this field chart ahead, with some notable preclinical advances in genetic forms of hearing loss through gene augmentation (Akil et al., 2012; Dulon et al., 2018; Isgrig et al., 2017; Pan et al., 2017), gene silencing (Shibata et al., 2016), and optoge-netic approaches to optimize the resolution of cochlear implants (Keppeler et al., 2018; Wrobel et al., 2018).
Gene therapy research in promising areas, such as olfaction and pain, is growing and in many cases also explores the utility of AAV, most often using local routes of administration (Guedon et al., 2015; Williams et al., 2017).
Lysosomal Storage Diseases
Biochemically, LSDs are characterized by the abnormal degradation of the lysosomal content and the pathological accumulation of undigested materials that leads to cellular damage and multi-systemic clinical manifestations. Most LSDs have a progressive neurodegenerative course that cannot be improved by enzyme replacement therapy as most enzymes do not cross the BBB (Platt et al., 2018), even though allogenic hematopoietic stem cell transplantation can provide some therapeutic benefit for patients. While those disorders taken individually are rare, the incidence of LSDs as a group is estimated as 1 in 5,000. Gene therapy applied to the treatment of LSDs is especially attractive for several reasons: (1) LSDs are monogenic diseases (recessive or X-linked),( 2) the natural history of those disorders is well established, (3) there is a therapeutic window during which an intervention can take place before overt neurodegeneration, (4) mutation carriers can easily been identified from families, (5) even low levels of residual enzymatic activity can result in significant clinical improvement, and (6) those diseases are often characterized by a severe phenotype and no alternative therapeutic options are available to correct the neurological deficits in patients. In addition, cross correction is made possible due to the fact that the lysosomal enzymes can be secreted in the extracellular space and taken up by surrounding cells, amplifying the corrective effect of the therapeutic gene.
Mucopolysaccharidoses (MPSs) encompass a group of rare metabolic diseases that arise from a deficit in the degradation of mucopolysaccharides and the resulting accumulation of gly-cosaminoglycans in the lysosome. The age of clinical onset directly correlates with the residual enzyme activity, which is why symptoms can be detected from the prenatal period through adulthood. CNS involvement encompasses difficulty of speech, ataxia, weakness, and dyskinesia in adults, while neurological impairment, developmental delay, and regression are observed in children whose life expectancy is significantly shortened (Platt et al., 2018). Several gene therapy clinical trials are currently in progress for the treatment of MPS I (Hurler infantile syndrome or Hurler-Scheie and Scheie for the juvenile and adult forms), MPS II (Hunter syndrome), MPS IIIA and MPS IIIB (Sanfilippo syndromes A and B), and MPS VI (Maroteaux-Lamy syndrome, which will not be discussed here as the central involvement is not a primary clinical symptom).
The MPS I (NCT02702115) and II (NCT03041324) phase I/II trials are of particular interest as they will both evaluate the safety and tolerability of ascending doses of rAAV2/6-based ZFN therapeutic agents to insert a correct functional copy of alpha-L-iduronidase (IDUA) or iduronate 2-sulfatase (IDS), respectively, into the albumin locus of patient hepatocytes after intravenous delivery. For those studies, adult MPS I and MPS II patients are being recruited, and it is expected that the benefit will mostly be on the peripheral symptoms of these diseases. Whether or not a similar approach can be proposed for younger patients remains to be seen. This would be of interest considering that ZFN-mediated genome editing also prevented the development of neurocognitive deficits in young MPS II mice, even after systemic administration (Laoharawee et al., 2018).
Because MPS IIIA is predominantly a CNS disorder, the phase II/III gene therapy proposed for this disease (‘‘AAVance’’) will directly target the CNS via stereotaxic infusion of AAVrh10-h.SGSH. The main goal is to deliver a functional copy of the human N-sulfoglucosamine sulfohydrolase (h.SGSH) cDNA in both hemispheres of the brain through MRI-guided neurosurgery. This study will enroll 20 patients of 6 months and older. A phase I/II clinical trial (NCT01474343) by the same team had been previously done on four young patients (6 months to 2 years) with the aim of expressing altogether the human SGSH and SUMF1 (sul-fatase-modifying factor, a catalytic activator if SGSH) cDNA after intracerebral administration of AAVrh.10. Overall safe, there was only an indication of cognitive benefit in the youngest subject, suggesting that the disease may have been too advanced in other patients for a measurable improvement (Tardieu et al., 2014). A parallel approach by intravenous delivery of AAV9.hSGSH is being developed by Abenoa Therapeutics (phase I/II ABO-102 trial, NCT02716246). Early-phase clinical trials for the treatment of MPS IIIB (NCT03300453 and NCT03315182) propose to restore expression of N-sulfoglucos-amine sulfohydrolase (NAGLU) in patients after intracerebral or intravenous delivery of AAV5 and AAV9, respectively. Both strategies have showed positive impact on the clinical symptoms in rodent and dog models (Ellinwood et al., 2011; Fu et al., 2011).
Neural ceroid lipofuscinoses or Batten disease are a group of neurodegenerative storage disorders. One type of Batten—late infantile neuronal ceroid lipofuscinosis (LINCL) — is a fatal childhood neurodegenerative LSD caused by mutation in the CLN2 gene encoding for the tripeptidyl peptidase-I (TPP-I) protein. The CNS involvement for this disease is significant (myoclonic epilepsy, loss of cognitive and motor function, degeneration of the retina leading to blindness, and early death; Donsante and Boulis, 2018). Currently, four AAV-based clinical trials for Batten disease are under way (three are active and not recruiting while one is recruiting). Pre-clinical safety and efficacy of AAV-based gene therapy for Batten disease has been demonstrated in rats and African green monkeys (Sondhi et al., 2012). The first phase I trial (NCT00151216) started in 2005 with a protocol of intracerebral injection of AAV2CUhCLN2. Ten CLN2 disease patients, with the severe (group A) or moderate (group B) form of the disease, were enrolled in the study. Unfortunately, one of the patients deceased 49 days after treatment from uncontrollable status epilepticus. While imaging studies did not detect a significant change in disease progression, neuropsychological evaluation demonstrated some retardation of disease progression relative to historic controls. It is possible that the use of AAV2, which spreads less than many newly characterized serotypes, did not lead to enough production of TPP-I in situ, therefore leading to little therapeutic impact (Worgall et al., 2008). A follow-up phase I study conducted by Weill Cornell (NCT01161576) administered an AAVRh.10CUhCLN2 by direct intraparenchymal stereotaxic injection. In this study, abnormal T2 hyperintensities detected by MRI around the sites of injection led to lowering the dose for the remaining subjects to be enrolled in protocol. Another parallel study with enlarged inclusion criteria (broader genotypes are being included) is being conducted by the same institution on a smaller number of patients (NCT01414985).
Hereditary Leukodystrophy: Canavan Disease
Canavan disease (CD) is an autosomal recessive leukodystrophies caused by loss-of-function mutations in the gene aspartoacylase (ASPA). ASPA is broadly expressed in various tissues but plays an essential role in the degradation of N-acetylaspar-tate (NAA), one of the most abundant metabolites in the brain (Rigotti et al., 2011; Tallan, 1957). If deficient, NAA abnormally builds up in the CNS, leading to NAA acidemia and aciduria as well as a broad range of deleterious effects such as dysmyeli-nation and spongiform degeneration in the CNS (Appu et al., 2017; Janson et al., 2006). During life, patients commonly present with head lag, macrocephaly, hypotonia, ataxia, inadequate visual tracking, epilepsy, and intellectual disabilities (Hoshino and Kubota, 2014), and many patients die before adolescence.
Preclinical studies in ASPA knockout rodent models demonstrated the remarkable efficacy of AAV expressing ASPA to resolve most of the biochemical and neuropathological hallmarks of the disease (Matalon et al., 2003; McPhee et al., 2005). The strong rationale for ASPA gene therapy has logically led to a first phase I/II dose-escalating clinical trial for CD in 2012 (Leone et al., 2012). Three groups of patients (of 3 months old to 8 years old) underwent AAV2.ASPA stereotaxic injection into the parenchyma at six sites. A control group did not receive any intervention. Adverse events within the first 90 days after treatment were related to the neurosurgery procedure and generally transient. Immune responses to AAV2 appeared limited. NAA concentrations were measured longitudinally by single-voxel 1H-MRS and showed a monotonic increase of NAA in all regions of interest in the brain of control subjects. By contrast, the pathological accumulation of NAA was decreased in AAV.ASPA-treated patients, which was significant in the periventricular and frontal regions (primary targets for the therapy). Even more interestingly, some patients enrolled in this study showed an arrest or even reversal of brain atrophy assessed by MRI. This was not the case for other subjects who already had a significant degree of atrophy before gene therapy, which makes the case for intervening as early as possible for this disease. Finally, modest but significant improvement in the motor functions (for example, rolling ability) and other clinical outcomes (in particular the level of alertness) was observed in AAV.ASPA-treated patients at 18 months, which was again less pronounced in the older subjects. Decreased clinical seizure frequency was observed in 11 of 13 patients. Overall, the CD clinical trial confirms the long-term safety of intracerebral stereotaxic infusion of AAV and shows evidence of positive radiographic and clinical changes, which is not 10 years out for some patients.
Neurotransmitter Disorder: Aromatic L-amino Acid Decarboxylase Deficiency
Aromatic L-amino acid decarboxylase (AADC) is a key enzyme involved in the metabolic biosynthesis of dopamine and serotonin, two monoamine neurotransmitters. The clinical manifestation of AADC deficiency includes hypotonia, hypokinesia, dystonia, oculogyric crisis, and autonomic dysfunction (Hwu et al., 2013). In the most severe form of the disease, children never reached developmental milestones and are unable to hold their head, sit, or stand. Life expectancy for those patients does not go above 5–6 years of age, as no cure is available. The first gene therapy trial for AADC deficiency (Hwu et al., 2012) was inspired by other clinical studies using AAV-based gene transfer in PD delivered into the putamen, which also aimed at restoring AADC expression and was well tolerated (Christine et al., 2009; Muramatsu et al., 2010). This first trial included four pediatric patients who received CT-guided stereotaxic injection of AAV2.CMV.hAADC in the putamen and whose motor performances were compared with untreated patients in the natural history control group. Impressively, the treated subjects gained weight, and their motor scores and cognitive functions gradually improved after gene transfer, even though all of them already had substantial impairment before intervention. All of them achieved head control and could sit with support. Results from a follow-up study including ten more subjects have been recently published, confirming the earlier findings (Chien et al., 2017) (NCT01395641). Currently, three clinical trials following a similar protocol (intra-putamen injection of AAV2.hAADC) are ongoing (NCT01395641, NCT02852213, and NCT02926066), with two of them actively recruiting new patients.
Idiopathic Age-Related Neurodegenerative Diseases
Gene therapy, conceptually straightforward for monogenic disorders, is more challenging for sporadic age-related neurodegenerative diseases. Because those diseases often result from the influence of many different pathogenic alleles with small individual impact, finding a good therapeutic target can be challenging despite the well-characterized neuropathological hallmarks. In addition, the lack of a clear genetic marker to identify patients and the relative frequency of mixed pathology often hindered the recruitment of a homogeneous set of patients in clinical trials. On the other hand, the high incidence of those idiopathic diseases facilitates the design of double-blind, randomized, and placebo-controlled studies including larger groups of patients (up to 60 for NCT00985517, a gene therapy trial for PD). Of note, in these types of studies, subjects in the placebo group are subjected to a sham injection protocol that does not include a brain cannula, injection, or vector infusion in order not to subject the individual to the risks associated with the procedure.
PD is a common neurodegenerative disorder characterized by bradykinesia, rigidity, tremor, and gait dysfunction. The main neuropathological hallmark of this disease is a loss of dopaminergic neurons in the substantia nigra pars compacta (SNc). The first line of defense to compensate for the loss of dopamine from the SNc is for patients to take L-dopa, which will be actively converted to dopamine by AADC. However, as the dopaminergic neurons continue degenerating, the effect of L-dopa weans off with time. Logically, one of the most straightforward application of gene therapy for PD is to restore expression of AADC by AAV transduction of non-degenerating striatal neurons (the site of action of dopamine). Pre-clinical results have proven that this strategy could increase the level of dopamine over a long period of time (up to 7 years in NHPs) (Bankiewicz et al., 2000; Bankiewicz et al., 2006; Sanchez-Pernaute et al., 2001), and the results of one phase I safety study have been published (NCT00229736) (Christine et al., 2009; Eberling et al., 2008). A total of ten patients were initially injected in two dose cohorts with AAV2.CMV. hAADC vector targeting four sites in the putamen. The intervention appeared to be well tolerated, even though postoperative MRI scan revealed the presence of hemorrhagic events along the canula trajectory. Nevertheless, improvement of the Unified Parkinson’s Disease Rating Scale (UPDRS) scores was reported at 6 months, which persisted for at least 2 years. Those remarkable results logically prompted the launch of other similar trials (either active or currently recruiting) with higher doses of vector (NCT02418598, NCT01973543, and NCT03065192) and larger cohorts of patients (up to 42 individuals for NCT03562494). The first randomized, placebo surgery controlled and doubleblinded phase 2 clinical trial for AAV.hAADC (VY-AADC02, Voyager Therapeutics) in advanced PD started in June 2018 (NCT03562494) (Table 1).
Alternatively, AAV-driven expression of neurturin is the focus of three other gene therapy trials for PD. Neurturin is a neurotrophic factor analog of glial-cell-derived neurotrophic factor (GDNF) that had been shown to improve dopaminergic activity in aged monkeys and be beneficial to animal models of PD (Gasmi et al., 2007; Herzog et al., 2007; Kordower et al., 2006; Kotzbauer et al., 1996). In this case, the therapeutic agent does not directly target a causative pathological molecular pathway but offers neurotrophic support to the vulnerable neuronal population degenerating in the disease. After having demonstrated the safety of neurturin gene transfer in a phase I study (NCT00252850) (Marks et al., 2008), results of the first double-blind, randomized, and sham-surgery-controlled gene therapy trial has been published in 2010 (NCT00400634) (Marks et al., 2010). Patients with idiopathic PD were randomly assigned (2:1) to receive bilateral stereotactic intraputaminal injection of AAV2.neurturin or sham procedure. Except for the neurosurgical team, patients and medical personnel were blind to the nature of the treatment. Assessment of motor functions at 18, but not at 12, months (UPDRS scale) did show some benefit of the AAV. neurturin over the sham-treated group. Post-mortem examination of the brains of two AAV.neurturin-treated individuals (Bartus et al., 2011) revealed that NRTN immunostaining covered, on average, 15% of the putamen, with sparse tyrosine hydroxylase induction in the striatum (no impact on the SNc was observed). Follow-up clinical trials include an open-labeled, one-arm study using convection-enhanced delivery of AAV2.GDNF (NCT01621581) (Johnston et al., 2009; Su et al., 2009) and asecond trial double-blind, randomized, placebo-controlled trial at higher dose (NCT00985517).
Late-onset AD is the lead cause of age-related dementia and affects more than 40% of individuals 85 years and older in Europe and the USA (Brookmeyer et al., 2007; Burns and Iliffe, 2009), and no less than 112 therapeutic agents are currently being tested in clinical trials (Cummings et al., 2018). After a first asymptomatic phase (during which amyloid accumulation is already detectable), cognitive complaints are reported by patients, followed by progressive loss of short-term memory, gradual deterioration of cognition, and loss of autonomy. From a neuropathological point of view, amyloid beta peptides aggregate and form extracellular neuritic plaques while the accumulation of abnormally phosphorylated tau protein triggers the development of intraneuronal neurofibrillary tangles. So far, none of the clinical trials have proven to significantly affect the course of the disease (with the exception of the phase II BAN2401 immunotherapy trial, NCT01767311, which may have a modest positive impact on the progression of cognitive deficits in AD patients), and only one attempt at gene therapy has been translated in patients.
This phase I trial, completed in 2014, was aimed at increasing the levels of nerve growth factor (NGF) in the brain of patients (NCT00087789) (Rafii et al., 2014). Similar to the neurturin approach for PD, this therapeutic agent essentially provides neurotrophic support. The rationale for choosing NGF was to prevent dysfunctional changes of basal cholinergic neurons, which are especially vulnerable to AD-associated stressors (Bartus et al., 1982; Hampel et al., 2018; Whitehouse et al., 1982). This approach by AAV-driven gene transfer (AAV2.NGF or CERE-110) also proposed an alternative to using cholinesterase inhibitors, which are associated with marginal improvement but non-negligible side effects. AAV.NGF was injected to the nucleus basalis of Meynert, the main source of cholinergic input of the neocortex, in ten patients with mild to moderate AD. With the exception of an event of postsurgical hygroma, the surgical procedure was well tolerated. At the 24-month time point, the rate of clinical deterioration of AAV-treated patients did not appear to be exacerbated as compared to the expected rate of decline determined from observational studies. However, potential efficacy could not be tested because of the small size of the cohort and the probable moderate impact of the treatment, even though a marginal stabilization in brain glucose metabolism was observed. Eventually, five patients died from the disease 3.3 to 6 years after treatment, and the post-mortem analysis of their brains showed evidence of NGF expression limited to the area in close proximity to the needle tract. The good safety and tolerability of this therapeutic agent has led to the design of a multicenter, double-blind, sham-surgery-controlled phase 2 clinical trial of AAV2.NGF in patients with mild to moderate AD started in 2009 (NCT00876863). As of July 2018 (24 months post AAV injection), no significant difference was detected between the treatment and placebo groups on clinical outcome measures or on selected AD biomarkers (Rafii et al., 2018). Ways for improvement include the recruitment of patients at an earlier stage of the disease (an MMSE between 17 to 26 is considered as ‘‘mild dementia’’) and the use of a more efficient vector serotype able to spread in larger area of the brain.
In July 2018, another pending phase I AAV-based clinical trial for AD seeks to assess the safety and toxicity of intracisternal delivery of AAVrh.10hAPOE2 in 15 patients homozygous for the epsilon-4 allele of the apolipoprotein E gene (APOE4/APOE4) (Rosenberg et al., 2018). APOE4 is by far the most significant genetic risk factor of sporadic AD (Corder et al., 1993; Strittmatter and Roses, 1996), while the epsilon-2 allele of APOE lowers this risk by half and even alleviates the severity of familial cases of AD (Corder et al., 1994; West et al., 1994). Because APOE is a secreted protein, significant amounts were also detected in the CSF, therefore possibly affecting a larger portion of the cerebral tissue than the areas primarily targeted by the injections. Whether or not those results will pan out in human AD patients and whether expression of APOE2 in neuronal cells may cause unexpected adverse events will be addressed in the phase I study.
Spinal Muscular Atrophy
SMA is a monogenic disorder caused by the loss of function of the gene encoding the survival motor neuron 1 (SMN1) protein, which leads to the progressive loss of lower MNs. SMA1 is the most severe form of the disease and a common genetic cause of death among infants, with a median age at symptom onset of 1.2 months and a very low rate of survival passed above 20 months (Finkel et al., 2014). Gene therapy application for type I SMA has recently received much attention from the scientific community for several reasons: first, it was the first time that a therapeutic gene was delivered via intravenous injection of an AAV9 harboring a scAAV vector genome (Figure 1C); second, the clinical benefit observed in patients was remarkable (NCT02122952) (Mendell et al., 2017). In the SMN1 gene augmentation approach, 15 infants received a high dose of scAAV9.CB.hSMN (AVX-101) intravenously (Figure 3). The study design included two dose cohorts. The high dose led to increases in serum aminotransferase, a sign of liver toxicity, in the first treated patient, which was managed by oral prednisolone. Subsequent subjects all received prophylactic prednisolone 24 h before surgery and continued for a month after the vector infusion. As of August 2017, all the children had survived and reached 20 months without the need for permanent mechanical ventilation. All subjects presented increased motor functions from baseline (based on the Children’s Hospital of Philadelphia Infant Test of Neuromuscular Disorders, or CHOP INTEND, scale), with greater improvements observed in the high-dose cohort who also had been treated at younger ages. Most of the children of this group achieved head control, sitting, rolling over, or speaking, milestones that otherwise would have never been reached according to the well-documented natural history for this disease (Finkel et al., 2014). This program subsequently was taken to phase 3 studies for SMA type I and follow on clinical studies for other types of SMA (NCT03306277, NCT03505099, NCT03461289, and NCT03381729), results of which are pending (Table 1). Originally developed by AveXis for SMA type I, a Biological License Application was submitted to the US FDA in November 2018 by Novartis, which acquired AveXis.
Conclusions
Alleviating symptoms long term or addressing the underlying etiology of disease is the goal in any area of medicine; however, for disorders of the nervous system, this often becomes a necessity. The limited access to therapeutic tissues like the brain, spinal cord, retina, or cochlea, either pharmacologically or surgically, begs for a lasting, if not permanent, approach to intervene in the disease process. Conceptually, gene therapy can achieve this, by either adding or eliminating a major genetic determinant in genetically defined disorders or, in common complex disease, by introducing a disease modifier in a lasting manner within the difficult-to-reach cell, tissue, or compartment. Over the past decades, increased insights in the challenges to clinical gene therapy, technological improvements that surmount these hurdles, and the advances in disease genetics, mechanism, diagnosis, and management have led to a series of milestones in clinical gene therapy. A growing body of data now supports the safety and, in an increasing number of cases, the efficacy of gene therapies across various modalities of genetic cargo, routes of administration, and target tissues. A common denominator in many of these studies is the use of AAV, a small viral gene transfer vector system with strong neuronal tropism and a desirable safety profile. The utility of AAV has been proven in dozens of clinical studies ranging from small-dose administrations for neurosensory disorders to large-dose systemic injections that achieve broad transduction of MNs in the spinal cord resulting in demonstrable and—at times—remarkable clinical benefit to patients.
While the opportunity is great for bringing gene therapy to inherited rare diseases and beyond, emerging from the clinic remains a challenge even after the compelling proof-of-concept studies. From the scientific perspective, particularly for common complex disorders, gene therapy, like any other drug modality, requires more viable molecular disease targets and modifiers. One prospect is that non-AAV biologics, which in their current form require repeat administration (e.g., antisense oligonucleotides or antibody-based treatments in IRD or AD), are ‘‘derisked’’ independently to then be incorporated for sustained delivery in an AAV formulation. From the technological perspective, notwithstanding the progress made, there is a need for further advances such as improved control over the specificity, kinetics, and potency of gene transfer and expression that limit toxicities, unlock new therapeutic target cells such as microglia, and improve on the efficacies already noted. The efficiency of AAV9 for CNS targeting, for example, even with scAAV technology, remains far from saturating and with limited neuronal targeting. Vector engineering efforts highlight the potential of AAV to do better in these and other respects. Particularly for the highdose systemic AAV applications, improvements are needed to maintain a healthy margin on a therapeutic window. While that window is primarily determined by a toxicity ceiling, in the context of gene therapy currently, manufacturability and cost can also be determinants. The cost associated with the production at scale of the doses needed to treat a single subject, let alone a broader patient population, can be staggering and in some scenario’s cost prohibitive for an otherwise viable gene therapy to move to clinical studies. Lastly, the ability to address the etiology of disease and the potential to induce a durable correction—the properties that make gene therapy so compelling, as Keeler noted in 1947 (Keeler, 1947)—are disruptive to traditional models in the pharmaceutical industry, creating uncertainty to a broader adoption of AAV gene therapy by clinicians, drug makers, and insurers.
Nonetheless, the clinical success in genetically defined diseases, such as IRD and SMA, has catalyzed an increasing activity and creativity to think through these hurdles. The future for AAV gene therapy to make a marked impact for patients suffering from disorders of the nervous system is therefore bright.
ACKNOWLEDGMENTS
The authors would like to thank Anna Maurer and Eric Zinn for the AAV capsid renderings in Figure 1 and Casey Maguire for valuable additional input about the biology of AAVs and their clinical application. This work was supported by grants from the Foundation Fighting Blindness, Research to Prevent Blindness, Lonza Houston, Oxford Biomedica, gifts from Giving/Grousbeck, the Butler Family Foundation, the Candyce Henwood Fund, the Ush2a Consortium (L.H.V.), and funding from the National Institutes of Health NIA (1R56AG057454–01 and 1R00AG047336–01A1) (E.H.).
Footnotes
DECLARATION OF INTERESTS
E.H. is a consultant to AZTherapies. L.H.V. holds equity in and chairs the Scientific Advisory Board of Akouos, a gene therapy company focused on hearing disorders. L.H.V. is also inventor and has a royalty interest in various AAV technologies, including AAV9 and AncAAVs, which have been licensed to several of the companies discussed in this Review. L.H.V. is a consultant to Nightstar Therapeutics, Cobalt, Lonza Houston, Exonics, and Selecta Biosciences and has received travel reimbursement from AveXis (now Novartis). L.H.V. receives research funding from Lonza Houston, Oxford Biomedica, Selecta Biosciences, and Solid Biosciences.
REFERENCES
- Akil O, Seal RP, Burke K, Wang C, Alemi A, During M, Edwards RH, and Lustig LR (2012). Restoration of hearing in the VGLUT3 knockout mouse using virally mediated gene therapy. Neuron 75, 283–293. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Amado D, Mingozzi F, Hui D, Bennicelli JL, Wei Z, Chen Y, Bote E, Grant RL, Golden JA, Narfstrom K, et al. (2010). Safety and efficacy of subretinal readministration of a viral vector in large animals to treat congenital blindness. Sci. Transl. Med 2, 21ra16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Appu AP, Moffett JR, Arun P, Moran S, Nambiar V, Krishnan JKS, Puthillathu N, and Namboodiri AMA (2017). Increasing N-acetylaspartate in the brain during postnatal myelination does not cause the CNS pathologies of Canavan disease. Front. Mol. Neurosci 10, 161. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Armbruster N, Lattanzi A, Jeavons M, Van Wittenberghe L, Gjata B, Marais T, Martin S, Vignaud A, Voit T, Mavilio F, et al. (2016). Efficacy and biodistribution analysis of intracerebroventricular administration of an optimized scAAV9-SMN1 vector in a mouse model of spinal muscular atrophy. Mol. Ther. Methods Clin. Dev 3, 16060. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Auricchio A, Kobinger G, Anand V, Hildinger M, O’Connor E, Maguire AM, Wilson JM, and Bennett J (2001). Exchange of surface proteins impacts on viral vector cellular specificity and transduction characteristics: the retina as a model. Hum. Mol. Genet 10, 3075–3081. [DOI] [PubMed] [Google Scholar]
- Azzouz M, Hottinger A, Paterna JC, Zurn AD, Aebischer P, and Büeler H (2000). Increased motoneuron survival and improved neuromuscular function in transgenic ALS mice after intraspinal injection of an adeno-associated virus encoding Bcl-2. Hum. Mol. Genet 9, 803–811. [DOI] [PubMed] [Google Scholar]
- Bankiewicz KS, Eberling JL, Kohutnicka M, Jagust W, Pivirotto P, Bringas J, Cunningham J, Budinger TF, and Harvey-White J (2000). Convection-enhanced delivery of AAV vector in parkinsonian monkeys; in vivo detection of gene expression and restoration of dopaminergic function using pro-drug approach. Exp. Neurol 164, 2–14. [DOI] [PubMed] [Google Scholar]
- Bankiewicz KS, Forsayeth J, Eberling JL, Sanchez-Pernaute R, Pivirotto P, Bringas J, Herscovitch P, Carson RE, Eckelman W, Reutter B, and Cunningham J (2006). Long-term clinical improvement in MPTP-lesioned primates after gene therapy with AAV-hAADC. Mol. Ther 14, 564–570. [DOI] [PubMed] [Google Scholar]
- Bartus RT, Dean RL 3rd, Beer B, and Lippa AS (1982). The cholinergic hypothesis of geriatric memory dysfunction. Science 217, 408–414. [DOI] [PubMed] [Google Scholar]
- Bartus RT, Herzog CD, Chu Y, Wilson A, Brown L, Siffert J, Johnson EM Jr., Olanow CW, Mufson EJ, and Kordower JH (2011). Bioactivity of AAV2-neurturin gene therapy (CERE-120): differences between Parkinson’s disease and nonhuman primate brains. Mov. Disord 26, 27–36. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bell P, Wang L, Lebherz C, Flieder DB, Bove MS, Wu D, Gao GP, Wilson JM, and Wivel NA (2005). No evidence for tumorigenesis of AAV vectors in a large-scale study in mice. Mol. Ther 12, 299–306. [DOI] [PubMed] [Google Scholar]
- Belur LR, Temme A, Podetz-Pedersen KM, Riedl M, Vulchanova L, Robinson N, Hanson LR, Kozarsky KF, Orchard PJ, Frey WH 2nd, et al. (2017). Intranasal adeno-associated virus mediated gene delivery and expression of human iduronidase in the central nervous system: a noninvasive and effective approach for prevention of neurologic disease in mucopolysaccharidosis type I. Hum. Gene Ther 28, 576–587. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Benkhelifa-Ziyyat S, Besse A, Roda M, Duque S, Astord S, Carcenac R, Marais T, and Barkats M (2013). Intramuscular scAAV9-SMN injection mediates widespread gene delivery to the spinal cord and decreases disease severity in SMA mice. Mol. Ther 21, 282–290. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Benraiss A, Toner MJ, Xu Q, Bruel-Jungerman E, Rogers EH, Wang F, Economides AN, Davidson BL, Kageyama R, Nedergaard M, and Goldman SA (2013). Sustained mobilization of endogenous neural progenitors delays disease progression in a transgenic model of Huntington’s disease. Cell Stem Cell 12, 787–799. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bey K, Ciron C, Dubreil L, Deniaud J, Ledevin M, Cristini J, Blouin V, Aubourg P, and Colle MA (2017). Efficient CNS targeting in adult mice by intrathecal infusion of single-stranded AAV9-GFP for gene therapy of neurological disorders. Gene Ther 24, 325–332. [DOI] [PubMed] [Google Scholar]
- Biferi MG, Cohen-Tannoudji M, Cappelletto A, Giroux B, Roda M, Astord S, Marais T, Bos C, Voit T, Ferry A, and Barkats M (2017). A new AAV10-U7-mediated gene therapy prolongs survival and restores function in an ALS mouse model. Mol. Ther 25, 2038–2052. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Biffi A, Montini E, Lorioli L, Cesani M, Fumagalli F, Plati T, Baldoli C, Martino S, Calabria A, Canale S, et al. (2013). Lentiviral hematopoietic stem cell gene therapy benefits metachromatic leukodystrophy. Science 341, 1233158. [DOI] [PubMed] [Google Scholar]
- Bilsland LG, Sahai E, Kelly G, Golding M, Greensmith L, and Schiavo G (2010). Deficits in axonal transport precede ALS symptoms in vivo. Proc. Natl. Acad. Sci. USA 107, 20523–20528. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Borel F, Kay MA, and Mueller C (2014). Recombinant AAV as a platform for translating the therapeutic potential of RNA interference. Mol. Ther 22, 692–701. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Botta S, Marrocco E, de Prisco N, Curion F, Renda M, Sofia M, Lupo M, Carissimo A, Bacci ML, Gesualdo C, et al. (2016). Rhodopsin targeted transcriptional silencing by DNA-binding. eLife 5, e12242. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Boulis NM, Noordmans AJ, Song DK, Imperiale MJ, Rubin A, Leone P, During M, and Feldman EL (2003). Adeno-associated viral vector gene expression in the adult rat spinal cord following remote vector delivery. Neurobiol. Dis 14, 535–541. [DOI] [PubMed] [Google Scholar]
- Broekman ML, Baek RC, Comer LA, Fernandez JL, Seyfried TN, and Sena-Esteves M (2007). Complete correction of enzymatic deficiency and neurochemistry in the GM1-gangliosidosis mouse brain by neonatal adenoassociated virus-mediated gene delivery. Mol. Ther 15, 30–37. [DOI] [PubMed] [Google Scholar]
- Brookmeyer R, Johnson E, Ziegler-Graham K, and Arrighi HM (2007). Forecasting the global burden of Alzheimer’s disease. Alzheimers Dement 3, 186–191. [DOI] [PubMed] [Google Scholar]
- Burns A, and Iliffe S (2009). Alzheimer’s disease. BMJ 338, b158. [DOI] [PubMed] [Google Scholar]
- Carson MJ, Doose JM, Melchior B, Schmid CD, and Ploix CC (2006). CNS immune privilege: hiding in plain sight. Immunol. Rev 213, 48–65. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cartier N, Hacein-Bey-Abina S, Bartholomae CC, Veres G, Schmidt M, Kutschera I, Vidaud M, Abel U, Dal-Cortivo L, Caccavelli L, et al. (2009). Hematopoietic stem cell gene therapy with a lentiviral vector in X-linked adrenoleukodystrophy. Science 326, 818–823. [DOI] [PubMed] [Google Scholar]
- Carty N, Lee D, Dickey C, Ceballos-Diaz C, Jansen-West K, Golde TE, Gordon MN, Morgan D, and Nash K (2010). Convection-enhanced delivery and systemic mannitol increase gene product distribution of AAV vectors 5, 8, and 9 and increase gene product in the adult mouse brain. J. Neurosci. Methods 194, 144–153. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Castle MJ, Gershenson ZT, Giles AR, Holzbaur EL, and Wolfe JH (2014). Adeno-associated virus serotypes 1, 8, and 9 share conserved mechanisms for anterograde and retrograde axonal transport. Hum. Gene Ther 25, 705–720. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Castle MJ, Turunen HT, Vandenberghe LH, and Wolfe JH (2016). Controlling AAV Tropism in the Nervous System with Natural and Engineered Capsids. Methods Mol. Biol 1382, 133–149. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cearley CN, and Wolfe JH (2007). A single injection of an adeno-associated virus vector into nuclei with divergent connections results in widespread vector distribution in the brain and global correction of a neurogenetic disease. J. Neurosci 27, 9928–9940. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chakrabarty P, Rosario A, Cruz P, Siemienski Z, Ceballos-Diaz C, Crosby K, Jansen K, Borchelt DR, Kim JY, Jankowsky JL, et al. (2013). Capsid serotype and timing of injection determines AAV transduction in the neonatal mice brain. PLoS ONE 8, e67680. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Challis RC, Ravindra Kumar S, Chan KY, Challis C, Beadle K, Jang MJ, Kim HM, Rajendran PS, Tompkins JD, Shivkumar K, et al. (2019). Systemic AAV vectors for widespread and targeted gene delivery in rodents. Nat. Protoc 14, 379–414. [DOI] [PubMed] [Google Scholar]
- Chandler RJ, LaFave MC, Varshney GK, Trivedi NS, Carrillo-Carrasco N, Senac JS, Wu W, Hoffmann V, Elkahloun AG, Burgess SM, and Venditti CP (2015). Vector design influences hepatic genotoxicity after adeno-associated virus gene therapy. J. Clin. Invest 125, 870–880. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chandler RJ, Sands MS, and Venditti CP (2017). Recombinant adeno-associated viral integration and genotoxicity: insights from animal models. Hum. Gene Ther 28, 314–322. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Charlesworth CT, Deshpande PS, Dever DP, Dejene B, Gomez-Ospina N, Mantri S, Pavel-Dinu M, Camarena J, Weinberg KI, and Porteus MH (2018). Identification of pre-existing adaptive immunity to Cas9 proteins in humans. bioRxiv https://doi.org/10.110½43345. [Google Scholar]
- Chauhan AN, and Lewis PD (1979). A quantitative study of cell proliferation in ependyma and choroid plexus in the postnatal rat brain. Neuropathol. Appl. Neurobiol 5, 303–309. [DOI] [PubMed] [Google Scholar]
- Chien YH, Lee NC, Tseng SH, Tai CH, Muramatsu SI, Byrne BJ, and Hwu WL (2017). Efficacy and safety of AAV2 gene therapy in children with aromatic L-amino acid decarboxylase deficiency: an open-label, phase ½ trial. Lancet Child Adolesc Health 1, 265–273. [DOI] [PubMed] [Google Scholar]
- Choudhury SR, Fitzpatrick Z, Harris AF, Maitland SA, Ferreira JS, Zhang Y, Ma S, Sharma RB, Gray-Edwards HL, Johnson JA, et al. (2016a). In vivo selection yields AAV-B1 capsid for central nervous system and muscle gene therapy. Mol. Ther 24, 1247–1257. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Choudhury SR, Harris AF, Cabral DJ, Keeler AM, Sapp E, Ferreira JS, Gray-Edwards HL, Johnson JA, Johnson AK, Su Q, et al. (2016b). Widespread central nervous system gene transfer and silencing after systemic delivery of novel AAV-AS vector. Mol. Ther 24, 726–735. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Choudhury SR, Hudry E, Maguire CA, Sena-Esteves M, Breakefield XO, and Grandi P (2017). Viral vectors for therapy of neurologic diseases. Neuropharmacology 120, 63–80. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Christine CW, Starr PA, Larson PS, Eberling JL, Jagust WJ, Hawkins RA, VanBrocklin HF, Wright JF, Bankiewicz KS, and Aminoff MJ (2009). Safety and tolerability of putaminal AADC gene therapy for Parkinson disease. Neurology 73, 1662–1669. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ciron C, Cressant A, Roux F, Raoul S, Cherel Y, Hantraye P, Deglon N, Schwartz B, Barkats M, Heard JM, et al. (2009). Human alpha-iduronidase gene transfer mediated by adeno-associated virus types 1,2, and 5 in the brain of nonhuman primates: vector diffusion and biodistribution. Hum. Gene Ther 20, 350–360. [DOI] [PubMed] [Google Scholar]
- Colle MA, Piguet F, Bertrand L, Raoul S, Bieche I, Dubreil L, Sloothaak D, Bouquet C, Moullier P, Aubourg P, et al. (2010). Efficient intracerebral delivery of AAV5 vector encoding human ARSA in non-human primate. Hum. Mol. Genet 19, 147–158. [DOI] [PubMed] [Google Scholar]
- Corder EH, Saunders AM, Strittmatter WJ, Schmechel DE, Gaskell PC, Small GW, Roses AD, Haines JL, and Pericak-Vance MA (1993). Gene dose of apolipoprotein E type 4 allele and the risk of Alzheimer’s disease in late onset families. Science 261, 921–923. [DOI] [PubMed] [Google Scholar]
- Corder EH, Saunders AM, Risch NJ, Strittmatter WJ, Schmechel DE, Gaskell PC Jr., Rimmler JB, Locke PA, Conneally PM, Schmader KE, et al. (1994). Protective effect of apolipoprotein E type 2 allele for late onset Alzheimer disease. Nat. Genet 7, 180–184. [DOI] [PubMed] [Google Scholar]
- Cukras C, Wiley HE, Jeffrey BG, Sen HN, Turriff A, Zeng Y, Vijayasarathy C, Marangoni D, Ziccardi L, Kjellstrom S, et al. (2018). Retinal AAV8-RS1 gene therapy for X-linked retinoschisis: initial findings from a phase I/IIa trial by intravitreal delivery. Mol. Ther 26, 2282–2294. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cummings J, Lee G, Ritter A, and Zhong K (2018). Alzheimer’s disease drug development pipeline: 2018. Alzheimers Dement (N Y) 4, 195–214. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dalkara D, Byrne LC, Klimczak RR, Visel M, Yin L, Merigan WH, Flannery JG, and Schaffer DV (2013). In vivo-directed evolution of a new adeno-associated virus for therapeutic outer retinal gene delivery from the vitreous. Sci. Transl. Med 5, 189ra76. [DOI] [PubMed] [Google Scholar]
- Dashkoff J, Lerner EP, Truong N, Klickstein JA, Fan Z, Mu D, Maguire CA, Hyman BT, and Hudry E (2016). Tailored transgene expression to specific cell types in the central nervous system after peripheral injection with AAV9. Mol. Ther. Methods Clin. Dev 3, 16081. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Davidson BL, Stein CS, Heth JA, Martins I, Kotin RM, Derksen TA, Zabner J, Ghodsi A, and Chiorini JA (2000). Recombinant adeno-associated virus type 2, 4, and 5 vectors: transduction of variant cell types and regions in the mammalian central nervous system. Proc. Natl. Acad. Sci. USA 97, 3428–3432. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Deverman BE, Pravdo PL, Simpson BP, Kumar SR, Chan KY, Banerjee A, Wu WL, Yang B, Huber N, Pasca SP, and Gradinaru V (2016). Cre-dependent selection yields AAV variants for widespread gene transfer to the adult brain. Nat. Biotechnol 34, 204–209. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Deverman BE, Ravina BM, Bankiewicz KS, Paul SM, and Sah DWY (2018). Gene therapy for neurological disorders: progress and prospects. Nat. Rev. Drug Discov 17, 641–659. [DOI] [PubMed] [Google Scholar]
- DiCarlo JE, Mahajan VB, and Tsang SH (2018). Gene therapy and genome surgery in the retina. J. Clin. Invest 128, 2177–2188. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dirren E, Aebischer J, Rochat C, Towne C, Schneider BL, and Aebischer P (2015). SOD1 silencing in motoneurons or glia rescues neuromuscular function in ALS mice. Ann. Clin. Transl. Neurol 2, 167–184. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dong B, Nakai H, and Xiao W (2010). Characterization of genome integrity for oversized recombinant AAV vector. Mol. Ther 18, 87–92. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Donsante A, and Boulis NM (2018). Progress in gene and cell therapies for the neuronal ceroid lipofuscinoses. Expert Opin. Biol. Ther 18, 755–764. [DOI] [PubMed] [Google Scholar]
- Donsante A, Miller DG, Li Y, Vogler C, Brunt EM, Russell DW, and Sands MS (2007). AAV vector integration sites in mouse hepatocellular carcinoma. Science 377, 477. [DOI] [PubMed] [Google Scholar]
- Duan D, Sharma P, Yang J, Yue Y, Dudus L, Zhang Y, Fisher KJ, and Engelhardt JF (1998). Circular intermediates of recombinant adeno-associated virus have defined structural characteristics responsible for long-term episomal persistence in muscle tissue. J. Virol 72, 8568–8577. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dufour BD, Smith CA, Clark RL, Walker TR, and McBride JL (2014). Intrajugular vein delivery of AAV9-RNAi prevents neuropathological changes and weight loss in Huntington’s disease mice. Mol. Ther 22, 797–810. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dulon D, Papal S, Patni P, Cortese M, Vincent PF, Tertrais M, Emptoz A, Tlili A, Bouleau Y, Michel V, et al. (2018). Clarin-1 gene transfer rescues auditory synaptopathy in model of Usher syndrome. J. Clin. Invest 128, 3382–3401. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Duque S, Joussemet B, Riviere C, Marais T, Dubreil L, Douar AM, Fyfe J, Moullier P, Colle MA, and Barkats M (2009). Intravenous administration of self-complementary AAV9 enables transgene delivery to adult motor neurons. Mol. Ther 17, 1187–1196. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Eberling JL, Jagust WJ, Christine CW, Starr P, Larson P, Bankiewicz KS, and Aminoff MJ (2008). Results from a phase I safety trial of hAADC gene therapy for Parkinson disease. Neurology 70, 1980–1983. [DOI] [PubMed] [Google Scholar]
- Eichler F, Duncan C, Musolino PL, Orchard PJ, De Oliveira S, Thrasher AJ, Armant M, Dansereau C, Lund TC, Miller WP, et al. (2017). Hematopoietic stem-cell gene therapy for cerebral adrenoleukodystrophy. N. Engl. J. Med 377, 1630–1638. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ellinwood NM, Ausseil J, Desmaris N, Bigou S, Liu S, Jens JK, Snella EM, Mohammed EE, Thomson CB, Raoul S, et al. (2011). Safe, efficient, and reproducible gene therapy of the brain in the dog models of Sanfilippoand Hurler syndromes. Mol. Ther 19, 251–259. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fiandaca MS, Varenika V, Eberling J, McKnight T, Bringas J, Pivirotto P, Beyer J, Hadaczek P, Bowers W, Park J, et al. (2009). Real-time MR imaging of adeno-associated viral vector delivery to the primate brain. Neuroimage 47 (Suppl 2), T27–T35. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Finkel RS, McDermott MP, Kaufmann P, Darras BT, Chung WK, Sproule DM, Kang PB, Foley AR, Yang ML, Martens WB, et al. (2014). Observational study of spinal muscular atrophy type I and implications for clinical trials. Neurology 83, 810–817. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fischer LR, Culver DG, Tennant P, Davis AA, Wang M, CastellanoSanchez A, Khan J, Polak MA, and Glass JD (2004). Amyotrophic lateral sclerosis is a distal axonopathy: evidence in mice and man. Exp. Neurol 185, 232–240. [DOI] [PubMed] [Google Scholar]
- Fisher KJ, Gao GP, Weitzman MD, DeMatteo R, Burda JF, and Wilson JM (1996). Transduction with recombinant adeno-associated virus for gene therapy is limited by leading-strand synthesis. J. Virol 70, 520–532. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Forrester JV, McMenamin PG, and Dando SJ (2018). CNS infection and immune privilege. Nat. Rev. Neurosci 19, 655–671. [DOI] [PubMed] [Google Scholar]
- Foust KD, Nurre E, Montgomery CL, Hernandez A, Chan CM, and Kaspar BK (2009). Intravascular AAV9 preferentially targets neonatal neurons and adult astrocytes. Nat. Biotechnol 27, 59–65. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Foust KD, Wang X, McGovern VL, Braun L, Bevan AK, Haidet AM, Le TT, Morales PR, Rich MM, Burghes AH, and Kaspar BK (2010). Rescue of the spinal muscular atrophy phenotype in a mouse model by early postnatal delivery of SMN. Nat. Biotechnol 28, 271–274. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
- Franz CK, Federici T, Yang J, Backus C, Oh SS, Teng Q, Carlton E, Bishop KM, Gasmi M, Bartus RT, et al. (2009). Intraspinal cord delivery of IGF-I mediated by adeno-associated virus 2 is neuroprotective in a rat model of familial ALS. Neurobiol. Dis 88, 473–481. [DOI] [PubMed] [Google Scholar]
- Friedmann T (1992). A brief history of gene therapy. Nat. Genet 2, 93–98. [DOI] [PubMed] [Google Scholar]
- Fu H, Dirosario J, Killedar S, Zaraspe K, and McCarty DM (2011). Correction of neurological disease of mucopolysaccharidosis IIIB in adult mice by rAAV9 trans-blood-brain barrier gene delivery. Mol. Ther 19, 1025–1033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gao G, Alvira MR, Somanathan S, Lu Y, Vandenberghe LH, Rux JJ, Calcedo R, Sanmiguel J, Abbas Z, and Wilson JM (2003). Adeno-associated viruses undergo substantial evolution in primates during natural infections. Proc. Natl. Acad. Sci. USA 100, 6081–6086. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gao G, Vandenberghe LH, and Wilson JM (2005). New recombinant serotypes of AAV vectors. Curr. Gene Ther 5, 285–297. [DOI] [PubMed] [Google Scholar]
- Garanto A, Chung DC, Duijkers L, Corral-Serrano JC, Messchaert M, Xiao R, Bennett J, Vandenberghe LH, and Collin RW (2016). In vitro and in vivo rescue of aberrant splicing in CEP290-associated LCA by antisense oligonucleotide delivery. Hum. Mol. Genet 25, 2552–2563. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gasmi M, Brandon EP, Herzog CD, Wilson A, Bishop KM, Hofer EK, Cunningham JJ, Printz MA, Kordower JH, and Bartus RT (2007). AAV2-mediated delivery of human neurturin to the rat nigrostriatal system: long-term efficacy and tolerability of CERE-120 for Parkinson’s disease. Neurobiol. Dis 27, 67–76. [DOI] [PubMed] [Google Scholar]
- Géléoc GS, and Holt JR (2014). Sound strategies for hearing restoration. Science 344, 1241062. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Georgiou E, Sidiropoulou K, Richter J, Papaneophytou C, Sargiannidou I, Kagiava A, von Jonquieres G, Christodoulou C, Klugmann M, and Kleopa KA (2017). Gene therapy targeting oligodendrocytes provides therapeutic benefit in a leukodystrophy model. Brain 140, 599–616. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gil-Farina I, Fronza R, Kaeppel C, Lopez-Franco E, Ferreira V, D’Avola D, Benito A, Prieto J, Petry H, Gonzalez-Aseguinolaza G, and Schmidt M (2016). Recombinant AAV integration is not associated with hepatic genotoxicity in nonhuman primates and patients. Mol. Ther 24, 1100–1105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ginn SL, Amaya AK, Alexander IE, Edelstein M, and Abedi MR (2018). Gene therapy clinical trials worldwide to 2017: an update. J. Gene Med 20, e3015. [DOI] [PubMed] [Google Scholar]
- Gray SJ, Matagne V, Bachaboina L, Yadav S, Ojeda SR, and Samulski RJ (2011). Preclinical differences of intravascular AAV9 delivery to neurons and glia: a comparative study of adult mice and nonhuman primates. Mol. Ther 19, 1058–1069. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gray SJ, Nagabhushan Kalburgi S, McCown TJ, and Jude Samulski R (2013). Global CNS gene delivery and evasion of anti–AAV-neutralizing antibodies by intrathecal AAV administration in non-human primates. Gene Ther 20, 450–459. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Grimm D, and Büning H (2017). Small but increasingly mighty: latest advances in AAV vector research, design, and evolution. Hum. Gene Ther 28, 1075–1086. [DOI] [PubMed] [Google Scholar]
- Grimm D, and Zolotukhin S (2015). E pluribus unum: 50 years of research, millions of viruses, and one goal-tailored acceleration of AAV evolution. Mol. Ther 23, 1819–1831. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guedon JM, Wu S, Zheng X, Churchill CC, Glorioso JC, Liu CH, Liu S, Vulchanova L, Bekker A, Tao YX, et al. (2015). Current gene therapy using viral vectors for chronic pain. Mol. Pain 11, 27. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gurda BL, De Guilhem De Lataillade A, Bell P, Zhu Y, Yu H, Wang P, Bagel J, Vite CH, Sikora T, Hinderer C, et al. (2016). Evaluation of AAV--mediated gene therapy for central nervous system disease in canine mucopolysaccharidosis VII. Mol. Ther 24, 206–216. [DOI] [PMC free article] [PubMed] [Google Scholar]
- György B, Sage C, Indzhykulian AA, Scheffer DI, Brisson AR, Tan S, Wu X, Volak A, Mu D, Tamvakologos PI, et al. (2017). Rescue of hearing by gene delivery to inner-ear hair cells using exosome-associated AAV. Mol. Ther 25, 379–391. [DOI] [PMC free article] [PubMed] [Google Scholar]
- György B, Meijer EJ, Ivanchenko MV, Tenneson K, Emond F, Hanlon KS, Indzhykulian AA, Volak A, Karavitaki KD, Tamvakologos PI, et al. (2018). Gene transfer with AAV9-PHP.B rescues hearing in a mouse model of usher syndrome 3A and transduces hair cells in a non-human primate. Mol. Ther. Methods Clin. Dev 13, 1–13. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hadaczek P, Kohutnicka M, Krauze MT, Bringas J, Pivirotto P, Cunningham J, and Bankiewicz K (2006). Convection-enhanced delivery of adeno-associated virus type 2 (AAV2) into the striatum and transport of AAV2 within monkey brain. Hum. Gene Ther 17, 291–302. [DOI] [PubMed] [Google Scholar]
- Hadaczek P, Eberling JL, Pivirotto P, Bringas J, Forsayeth J, and Bankiewicz KS (2010). Eight years of clinical improvement in MPTP-lesioned primates after gene therapy with AAV2-hAADC. Mol. Ther 18, 1458–1461. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hampel H, Mesulam MM, Cuello AC, Farlow MR, Giacobini E, Grossberg GT, Khachaturian AS, Vergallo A, Cavedo E, Snyder PJ, and Khachaturian ZS (2018). The cholinergic system in the pathophysiology and treatment of Alzheimer’s disease. Brain 141, 1917–1933. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Han CL, Ge M, Liu YP, Zhao XM, Wang KL, Chen N, Meng WJ, Hu W, Zhang JG, Li L, and Meng FG (2018). LncRNA H19 contributes to hippocampal glial cell activation via JAK/STAT signaling in a rat model of temporal lobe epilepsy. J. Neuroinflammation 15, 103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hardcastle N, Boulis NM, and Federici T (2018). AAV gene delivery to the spinal cord: serotypes, methods, candidate diseases, and clinical trials. Expert Opin. Biol. Ther 18, 293–307. [DOI] [PubMed] [Google Scholar]
- Harmatz P, Muenzer J, Burton BK, Ficicioglu C, Lau HA, Leslie ND, Conner E, Wong Po Foo C, Vaidya SA, Wechsler T, and Whitley CB (2018). Update on phase ½ clinical trials for MPS I and MPS II using ZFN-mediated in vivo genome editing. Mol. Genet. Metab 123, S59–S60. [Google Scholar]
- Hastie E, and Samulski RJ (2015). Adeno-associated virus at 50: a golden anniversary of discovery, research, and gene therapy success-a personal perspective. Hum. Gene Ther 26, 257–265. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Haurigot V, Marco S, Ribera A, Garcia M, Ruzo A, Villacampa P, Ayuso E, Anor S, Andaluz A, Pineda M, et al. (2013). Whole body correction of mucopolysaccharidosis IIIA by intracerebrospinal fluid gene therapy. J. Clin. Invest Published online July 1, 2013. 10.1172/JCI66778. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Herzog CD, Dass B, Holden JE, Stansell J 3rd, Gasmi M, Tuszynski MH, Bartus RT, and Kordower JH (2007). Striatal delivery of CERE-120, an AAV2 vector encoding human neurturin, enhances activity of the dopaminergic nigrostriatal system in aged monkeys. Mov. Disord 22, 1124–1132. [DOI] [PubMed] [Google Scholar]
- Hinderer C, Katz N, Louboutin JP, Bell P, Yu H, Nayal M, Kozarsky K, O’Brien WT, Goode T, and Wilson JM (2016). Delivery of an adeno-associated virus vector into cerebrospinal fluid attenuates central nervous system disease in mucopolysaccharidosis type II mice. Hum. Gene Ther 27,906–915. [DOI] [PubMed] [Google Scholar]
- Hinderer C, Bell P, Katz N, Vite CH, Louboutin JP, Bote E, Yu H, Zhu Y, Casal ML, Bagel J, et al. (2018a). Evaluation of intrathecal routes of administration for adeno-associated viral vectors in large animals. Hum. Gene Ther 29, 15–24. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hinderer C, Katz N, Buza EL, Dyer C, Goode T, Bell P, Richman LK, and Wilson JM (2018b). Severe toxicity in nonhuman primates and piglets following high-dose intravenous administration of an adeno-associated virus vector expressing human SMN. Hum. Gene Ther 29, 285–298. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hironaka K, Yamazaki Y, Hirai Y, Yamamoto M, Miyake N, Miyake K, Okada T, Morita A, and Shimada T (2015). Enzyme replacement in the CSF to treat metachromatic leukodystrophy in mouse model using single intracerebroventricular injection of self-complementary AAV1 vector. Sci. Rep 5, 13104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hocquemiller M, Giersch L, Audrain M, Parker S, and Cartier N (2016). Adeno-associated virus-based gene therapy for CNS diseases. Hum. Gene Ther 27, 478–496. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hollis Ii ER, Kadoya K, Hirsch M, Samulski RJ, and Tuszynski MH (2008). Efficient retrograde neuronal transduction utilizing self-complementary AAV1. Mol. Ther 16, 296–301. [DOI] [PubMed] [Google Scholar]
- Holt JR, and Vandenberghe LH (2012). Gene therapy for deaf mice goes viral. Mol. Ther 20, 1836–1837. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hordeaux J, Dubreil L, Robveille C, Deniaud J, Pascal Q, Dequeant B, Pailloux J, Lagalice L, Ledevin M, Babarit C, et al. (2017). Long-term neurologic and cardiac correction by intrathecal gene therapy in Pompe disease. Acta Neuropathol. Commun 5, 66. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hordeaux J, Wang Q, Katz N, Buza EL, Bell P, and Wilson JM (2018). The neurotropic properties of AAV-PHP.B are limited to C57BL/6J mice. Mol. Ther 26, 664–668. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hoshino H, and Kubota M (2014). Canavan disease: clinical features and recent advances in research. Pediatr. Int 56, 477–483. [DOI] [PubMed] [Google Scholar]
- Hudry E, Dashkoff J, Roe AD, Takeda S, Koffie RM, Hashimoto T, Scheel M, Spires-Jones T, Arbel-Ornath M, Betensky R, et al. (2013). Gene transfer of human Apoe isoforms results in differential modulation of amyloid deposition and neurotoxicity in mouse brain. Sci. Transl. Med 5, 212ra161. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hudry E, Martin C, Gandhi S, Gyorgy B, Scheffer DI, Mu D, Merkel SF, Mingozzi F, Fitzpatrick Z, Dimant H, et al. (2016). Exosome-associated AAV vector as a robust and convenient neuroscience tool. Gene Ther 23, 380–392. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hudry E, Andres-Mateos E, Lerner EP, Volak A, Cohen O, Hyman BT, Maguire CA, and Vandenberghe LH (2018). Efficient gene transfer to the central nervous system by single-stranded Anc80L65. Mol. Ther. Methods Clin. Dev 10, 197–209. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hwu WL, Muramatsu S, Tseng SH, Tzen KY, Lee NC, Chien YH, Snyder RO, Byrne BJ, Tai CH, and Wu RM (2012). Gene therapy for aromatic L-amino acid decarboxylase deficiency. Sci. Transl. Med 4, 134ra61. [DOI] [PubMed] [Google Scholar]
- Hwu WL, Lee NC, Chien YH, Muramatsu S, and Ichinose H (2013). AADC deficiency: occurring in humans, modeled in rodents. Adv. Pharmacol 68, 273–284. [DOI] [PubMed] [Google Scholar]
- Isgrig K, Shteamer JW, Belyantseva IA, Drummond MC, Fitzgerald TS, Vijayakumar S, Jones SM, Griffith AJ, Friedman TB, Cunningham LL, and Chien WW (2017). Gene therapy restores balance and auditory functions in a mouse model of usher syndrome. Mol. Ther 25, 780–791. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Iwata N, Sekiguchi M, Hattori Y, Takahashi A, Asai M, Ji B, Higuchi M, Staufenbiel M, Muramatsu S, and Saido TC (2013). Global brain delivery of neprilysin gene by intravascular administration of AAV vector in mice. Sci. Rep 3, 1472. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jackson KL, Dayton RD, Deverman BE, and Klein RL (2016). Better targeting, better efficiency for wide-scale neuronal transduction with the synapsin promoter and AAV-PHP.B. Front. Mol. Neurosci 9, 116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Janson C, McPhee S, Bilaniuk L, Haselgrove J, Testaiuti M, Freese A, Wang DJ, Shera D, Hurh P, Rupin J, et al. (2002). Clinical protocol. Gene therapy of Canavan disease: AAV-2 vector for neurosurgical delivery of aspartoacylase gene (ASPA) to the human brain. Hum. Gene Ther 13, 1391–1412. [DOI] [PubMed] [Google Scholar]
- Janson CG, McPhee SW, Francis J, Shera D, Assadi M, Freese A, Hurh P, Haselgrove J, Wang DJ, Bilaniuk L, and Leone P (2006). Natural history of Canavan disease revealed by proton magnetic resonance spectroscopy (1H-MRS) and diffusion-weighted MRI. Neuropediatrics 37, 209–221. [DOI] [PubMed] [Google Scholar]
- Jergova S, Gordon CE, Gajavelli S, and Sagen J (2017). Experimental gene therapy with serine-histogranin and endomorphin 1 for the treatment of chronic neuropathic pain. Front. Mol. Neurosci 10, 406. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Johnston LC, Eberling J, Pivirotto P, Hadaczek P, Federoff HJ, Forsayeth J, and Bankiewicz KS (2009). Clinically relevant effects of convection-enhanced delivery of AAV2-GDNF on the dopaminergic nigrostriatal pathway in aged rhesus monkeys. Hum. Gene Ther 20, 497–510. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kanaan NM, Sellnow RC, Boye SL, Coberly B, Bennett A, Agbandje-McKenna M, Sortwell CE, Hauswirth WW, Boye SE, and Manfredsson FP (2017). Rationally engineered AAV capsids improve transduction and volumetric spread in the CNS. Mol. Ther. Nucleic Acids 8, 184–197. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Karali M, Manfredi A, Puppo A, Marrocco E, Gargiulo A, Allocca M, Corte MD, Rossi S, Giunti M, Bacci ML, et al. (2011). MicroRNA-restricted transgene expression in the retina. PLoS ONE 6, e22166. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Karumuthil-Melethil S, Marshall MS, Heindel C, Jakubauskas B, Bongarzone ER, and Gray SJ (2016). Intrathecal administration of AAV/GALC vectors in 10–11-day-old twitcher mice improves survival and is enhanced by bone marrow transplant. J. Neurosci. Res 94, 1138–1151. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Keeler CE (1947). Gene therapy. J. Hered 38, 294–298. [PubMed] [Google Scholar]
- Kennedy LH, and Rinholm JE (2017). Visualization and live imaging of oligodendrocyte organelles in organotypic brain slices using adeno-associated virus and confocal microscopy. J. Vis. Exp Published online October 23, 2017. 10.3791/56237. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Keppeler D, Merino RM, Lopez de la Morena D, Bali B, Huet AT, Gehrt A, Wrobel C, Subramanian S, Dombrowski T, Wolf F, et al. (2018). Ultrafast optogenetic stimulation of the auditory pathway by targeting-optimized Chronos. EMBO J 37, 37. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Khabou H, Cordeau C, Pacot L, Fisson S, and Dalkara D (2018). Dosage thresholds and influence of transgene cassette in adeno-associated virus-related toxicity. Hum. Gene Ther 29, 1235–1241. [DOI] [PubMed] [Google Scholar]
- Knipe DM, and Howley PM (2013). Fields Virology, Sixth Edition (Wolters Kluwer/Lippincott Williams & Wilkins Health; ). [Google Scholar]
- Koh W, Park YM, Lee SE, and Lee CJ (2017). AAV-mediated astrocytespecific gene expression under human ALDH1L1 promoter in mouse thalamus. Exp. Neurobiol 26, 350–361. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kordower JH, Herzog CD, Dass B, Bakay RA, Stansell J 3rd, Gasmi M, and Bartus RT (2006). Delivery of neurturin by AAV2 (CERE-120)-mediated gene transfer provides structural and functional neuroprotection and neurorestoration in MPTP-treated monkeys. Ann. Neurol 60, 706–715. [DOI] [PubMed] [Google Scholar]
- Kotin RM, Siniscalco M, Samulski RJ, Zhu XD, Hunter L, Laughlin CA, McLaughlin S, Muzyczka N, Rocchi M, and Berns KI (1990). Sitespecific integration by adeno-associated virus. Proc. Natl. Acad. Sci. USA 87, 2211–2215. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kotzbauer PT, Lampe PA, Heuckeroth RO, Golden JP, Creedon DJ, Johnson EM Jr., and Milbrandt J (1996). Neurturin, a relative of glial-cellline-derived neurotrophic factor. Nature 384, 467–470. [DOI] [PubMed] [Google Scholar]
- Lai Y, Yue Y, and Duan D (2010). Evidence for the failure of adeno-associated virus serotype 5 to package a viral genome > or = 8.2 kb. Mol. Ther 18, 75–79. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Landegger LD, Pan B, Askew C, Wassmer SJ, Gluck SD, Galvin A, Taylor R, Forge A, Stankovic KM, Holt JR, and Vandenberghe LH (2017). A synthetic AAV vector enables safe and efficient gene transfer to the mammalian inner ear. Nat. Biotechnol 35, 280–284. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Laoharawee K, DeKelver RC, Podetz-Pedersen KM, Rohde M, Sproul S, Nguyen HO, Nguyen T, St Martin SJ, Ou L, Tom S, et al. (2018). Dose-dependent prevention of metabolic and neurologic disease in murine MPS II by ZFN-mediated in vivo genome editing. Mol. Ther 26, 1127–1136. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Leinekugel P, Michel S, Conzelmann E, and Sandhoff K (1992). Quantitative correlation between the residual activity of beta-hexosaminidase A and arylsulfatase A and the severity of the resulting lysosomal storage disease. Hum. Genet 88, 513–523. [DOI] [PubMed] [Google Scholar]
- Leone P, Shera D, McPhee SW, Francis JS, Kolodny EH, Bilaniuk LT, Wang DJ, Assadi M, Goldfarb O, Goldman HW, et al. (2012). Long-term follow-up after gene therapy for canavan disease. Sci. Transl. Med 4, 165ra163. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Levites Y, Jansen K, Smithson LA, Dakin R, Holloway VM, Das P, and Golde TE (2006). Intracranial adeno-associated virus-mediated delivery of anti-pan amyloid beta, amyloid beta40, and amyloid beta42 single-chain variable fragments attenuates plaque pathology in amyloid precursor protein mice. J. Neurosci 26, 11923–11928. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li H, Malani N, Hamilton SR, Schlachterman A, Bussadori G, Edmonson SE, Shah R, Arruda VR, Mingozzi F, Wright JF, et al. (2011). Assessing the potential for AAV vector genotoxicity in a murine model. Blood 117, 3311–3319. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li D, Liu C, Yang C, Wang D, Wu D, Qi Y, Su Q, Gao G, Xu Z, and Guo Y (2017). Slow intrathecal injection of rAAVrh10 enhances its transduction of spinal cord and therapeutic efficacy in a mutant SOD1 model of ALS. Neuroscience 365, 192–205. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Linden RM, Ward P, Giraud C, Winocour E, and Berns KI (1996). Sitespecific integration by adeno-associated virus. Proc. Natl. Acad. Sci. USA 93, 11288–11294. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu G, Martins I, Wemmie JA, Chiorini JA, and Davidson BL (2005). Functional correction of CNS phenotypes in a lysosomal storage disease model using adeno-associated virus type 4 vectors. J. Neurosci 25, 9321–9327. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Logan GJ, Dane AP, Hallwirth CV, Smyth CM, Wilkie EE, Amaya AK, Zhu E, Khandekar N, Ginn SL, Liao SHY, et al. (2017). Identification of liver-specific enhancer-promoter activity in the 30 untranslated region of the wild-type AAV2 genome. Nat. Genet 49, 1267–1273. [DOI] [PubMed] [Google Scholar]
- Low K, Aebischer P, and Schneider BL (2013). Direct and retrograde transduction of nigral neurons with AAV6, 8, and 9 and intraneuronal persistence of viral particles. Hum. Gene Ther 24, 613–629. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lukason M, DuFresne E, Rubin H, Pechan P, Li Q, Kim I, Kiss S, Flaxel C, Collins M, Miller J, et al. (2011). Inhibition of choroidal neovascularization in a nonhuman primate model by intravitreal administration of an AAV2 vector expressing a novel anti-VEGF molecule. Mol. Ther 19, 260–265. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ma XC, Liu P, Zhang XL, Jiang WH, Jia M, Wang CX, Dong YY, Dang YH, and Gao CG (2016). Intranasal delivery of recombinant AAV containing BDNF fused with HA2TAT: a potential promising therapy strategy for major depressive disorder. Sci. Rep 6, 22404. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Maclachlan TK, Lukason M, Collins M, Munger R, Isenberger E, Rogers C, Malatos S, Dufresne E, Morris J, Calcedo R, et al. (2011). Preclinical safety evaluation of AAV2-sFLT01-a gene therapy for age-related macular degeneration. Mol. Ther 19, 326–334. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Maguire AM, Simonelli F, Pierce EA, Pugh EN Jr., Mingozzi F, Bennicelli J, Banfi S, Marshall KA, Testa F, Surace EM, et al. (2008). Safety and efficacy of gene transfer for Leber’s congenital amaurosis. N. Engl. J. Med 358, 2240–2248. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Manno CS, Pierce GF, Arruda VR, Glader B, Ragni M, Rasko JJ, Ozelo MC, Hoots K, Blatt P, Konkle B, et al. (2006). Successful transduction of liver in hemophilia by AAV-Factor IX and limitations imposed by the host immune response. Nat. Med 12, 342–347. [DOI] [PubMed] [Google Scholar]
- Marks WJ Jr., Ostrem JL, Verhagen L, Starr PA, Larson PS, Bakay RA, Taylor R, Cahn-Weiner DA, Stoessl AJ, Olanow CW, and Bartus RT (2008). Safety and tolerability of intraputaminal delivery of CERE-120 (adeno-associated virus serotype 2-neurturin) to patients with idiopathic Parkinson’s disease: an open-label, phase I trial. Lancet Neurol 7, 400–408. [DOI] [PubMed] [Google Scholar]
- Marks WJ Jr., Bartus RT, Siffert J, Davis CS, Lozano A, Boulis N, Vitek J, Stacy M, Turner D, Verhagen L, et al. (2010). Gene delivery of AAV2-neurturin for Parkinson’s disease: a double-blind, randomised, controlled trial. Lancet Neurol 9, 1164–1172. [DOI] [PubMed] [Google Scholar]
- Mastakov MY, Baer K, Xu R, Fitzsimons H, and During MJ (2001). Combined injection of rAAV with mannitol enhances gene expression in the rat brain. Mol. Ther 3, 225–232. [DOI] [PubMed] [Google Scholar]
- Mastakov MY, Baer K, Kotin RM, and During MJ (2002). Recombinant adeno-associated virus serotypes 2-and 5-mediated gene transfer in the mammalian brain: quantitative analysis of heparin co-infusion. Mol. Ther 5, 371–380. [DOI] [PubMed] [Google Scholar]
- Matalon R, Surendran S, Rady PL, Quast MJ, Campbell GA, Matalon KM, Tyring SK, Wei J, Peden CS, Ezell EL, et al. (2003). Adeno-associated virus-mediated aspartoacylase gene transfer to the brain of knockout mouse for canavan disease. Mol. Ther 7, 580–587. [DOI] [PubMed] [Google Scholar]
- Mattar CN, Waddington SN, Biswas A, Johana N, Ng XW, Fisk AS, Fisk NM, Tan LG, Rahim AA, Buckley SM, et al. (2013). Systemic delivery of scAAV9 in fetal macaques facilitates neuronal transduction of the central and peripheral nervous systems. Gene Ther 20, 69–83. [DOI] [PubMed] [Google Scholar]
- McBride JL, Pitzer MR, Boudreau RL, Dufour B, Hobbs T, Ojeda SR, and Davidson BL (2011). Preclinical safety of RNAi-mediated HTT suppression in the rhesus macaque as a potential therapy for Huntington’s disease. Mol.Ther 19, 2152–2162. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McCarty DM (2008). Self-complementary AAV vectors; advances and applications. Mol. Ther 16, 1648–1656. [DOI] [PubMed] [Google Scholar]
- McCarty DM, DiRosario J, Gulaid K, Muenzer J, and Fu H (2009). Mannitol-facilitated CNS entry of rAAV2 vector significantly delayed the neurological disease progression in MPS IIIB mice. Gene Ther 16, 1340–1352. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McCurdy VJ, Rockwell HE, Arthur JR, Bradbury AM, Johnson AK, Randle AN, Brunson BL, Hwang M, Gray-Edwards HL, Morrison NE, et al. (2015). Widespread correction of central nervous system disease after intracranial gene therapy in a feline model of Sandhoff disease. Gene Ther 22, 181–189. [DOI] [PubMed] [Google Scholar]
- McLean JR, Smith GA, Rocha EM, Hayes MA, Beagan JA, Hallett PJ, and Isacson O (2014). Widespread neuron-specific transgene expression in brain and spinal cord following synapsin promoter-driven AAV9 neonatal intracerebroventricular injection. Neurosci. Lett 576, 73–78. [DOI] [PubMed] [Google Scholar]
- McPhee SW, Francis J, Janson CG, Serikawa T, Hyland K, Ong EO, Raghavan SS, Freese A, and Leone P (2005). Effects of AAV-2-mediated aspartoacylase gene transfer in the tremor rat model of Canavan disease. Brain Res. Mol. Brain Res 135, 112–121. [DOI] [PubMed] [Google Scholar]
- Mendell JR, Al-Zaidy S, Shell R, Arnold WD, Rodino-Klapac LR, Prior TW, Lowes L, Alfano L, Berry K, Church K, et al. (2017). Single-dose gene-replacement therapy for spinal muscular atrophy. N. Engl. J. Med 377, 1713–1722. [DOI] [PubMed] [Google Scholar]
- Meyer K, Ferraiuolo L, Schmelzer L, Braun L, McGovern V, Likhite S, Michels O, Govoni A, Fitzgerald J, Morales P, et al. (2015). Improving single injection CSF delivery of AAV9-mediated gene therapy for SMA: a doseresponse study in mice and nonhuman primates. Mol. Ther 23, 477–487. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Miller AD (1992). Human gene therapy comes of age. Nature 357, 455–460. [DOI] [PubMed] [Google Scholar]
- Miller JW, and Vandenberghe LH (2018). Breaking and sealing barriers in retinal gene therapy. Mol. Ther 26, 2081–2082. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mingozzi F, and High KA (2013). Immune responses to AAV vectors: overcoming barriers to successful gene therapy. Blood 122, 23–36. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mingozzi F, Maus MV, Hui DJ, Sabatino DE, Murphy SL, Rasko JE, Ragni MV, Manno CS, Sommer J, Jiang H, et al. (2007). CD8(+) T-cell responses to adeno-associated virus capsid in humans. Nat. Med 13, 419–422. [DOI] [PubMed] [Google Scholar]
- Muramatsu S, Fujimoto K, Kato S, Mizukami H, Asari S, Ikeguchi K, Kawakami T, Urabe M, Kume A, Sato T, et al. (2010). A phase I study of aromatic L-amino acid decarboxylase gene therapy for Parkinson’s disease. Mol. Ther 18, 1731–1735. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Murrey DA, Naughton BJ, Duncan FJ, Meadows AS, Ware TA, Campbell KJ, Bremer WG, Walker CM, Goodchild L, Bolon B, et al. (2014). Feasibility and safety of systemic rAAV9-hNAGLU delivery for treating mucopolysaccharidosis IIIB: toxicology, biodistribution, and immunological assessments in primates. Hum. Gene Ther. Clin. Dev 25, 72–84. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nakai H, Montini E, Fuess S, Storm TA, Grompe M, and Kay MA (2003). AAV serotype 2 vectors preferentially integrate into active genes in mice. Nat. Genet 34, 297–302. [DOI] [PubMed] [Google Scholar]
- Nathwani AC, Tuddenham EG, Rangarajan S, Rosales C, McIntosh J, Linch DC, Chowdary P, Riddell A, Pie AJ, Harrington C, et al. (2011). Adenovirus-associated virus vector-mediated gene transfer in hemophilia B. N. Engl. J. Med 365, 2357–2365. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nault JC, Datta S, Imbeaud S, Franconi A, Mallet M, Couchy G, Letouze E, Pilati C, Verret B, Blanc JF, et al. (2015). Recurrent AAV2-related insertional mutagenesis in human hepatocellular carcinomas. Nat. Genet 47, 1187–1193. [DOI] [PubMed] [Google Scholar]
- Nault JC, Datta S, Imbeaud S, Franconi A, and Zucman-Rossi J (2016). Adeno-associated virus type 2 as an oncogenic virus in human hepatocellular carcinoma. Mol. Cell. Oncol 3, e1095271. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pan B, Askew C, Galvin A, Heman-Ackah S, Asai Y, Indzhykulian AA, Jodelka FM, Hastings ML, Lentz JJ, Vandenberghe LH, et al. (2017). Gene therapy restores auditory and vestibular function in a mouse model of Usher syndrome type 1c. Nat. Biotechnol 35, 264–272. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Passini MA, Watson DJ, Vite CH, Landsburg DJ, Feigenbaum AL, and Wolfe JH (2003). Intraventricular brain injection of adeno-associated virus type 1 (AAV1) in neonatal mice results in complementary patterns of neuronal transduction to AAV2 and total long-term correction of storage lesions in the brains of beta-glucuronidase-deficient mice. J. Virol 77, 7034–7040. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Peden MC, Min J, Meyers C, Lukowski Z, Li Q, Boye SL, Levine M, Hauswirth WW, Ratnakaram R, Dawson W, et al. (2011). Ab-externo AAV-mediated gene delivery to the suprachoroidal space using a 250 micron flexible microcatheter. PLoS ONE 6, e17140. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pillay S, and Carette JE (2017). Host determinants of adeno-associated viral vector entry. Curr. Opin. Virol 24, 124–131. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Platt RJ, Chen S, Zhou Y, Yim MJ, Swiech L, Kempton HR, Dahlman JE, Parnas O, Eisenhaure TM, Jovanovic M, et al. (2014). CRISPR-Cas9 knockin mice for genome editing and cancer modeling. Cell 159, 440–455. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Platt FM, d’Azzo A, Davidson BL, Neufeld EF, and Tifft CJ (2018). Lysosomal storage diseases. Nat. Rev. Dis. Primers 4, 27. [DOI] [PubMed] [Google Scholar]
- Pleger ST, Shan C, Ksienzyk J, Bekeredjian R, Boekstegers P, Hinkel R, Schinkel S, Leuchs B, Ludwig J, Qiu G, et al. (2011). Cardiac AAV9-S100A1 gene therapy rescues post-ischemic heart failure in a preclinical large animal model. Sci. Transl. Med 3, 92ra64. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rafii MS, Baumann TL, Bakay RA, Ostrove JM, Siffert J, Fleisher AS, Herzog CD, Barba D, Pay M, Salmon DP, et al. (2014). A phase1 study of stereotactic gene delivery of AAV2-NGF for Alzheimer’s disease. Alzheimers Dement 10, 571–581. [DOI] [PubMed] [Google Scholar]
- Rafii MS, Tuszynski MH, Thomas RG, Barba D, Brewer JB, Rissman RA, Siffert J, and Aisen PS; AAV2-NGF Study Team (2018). Adeno-associated viral vector (serotype 2)-nerve growth factor for patients with Alzheimer disease: a randomized clinical trial. JAMA Neurol 75, 834–841. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Reichel FF, Dauletbekov DL, Klein R, Peters T, Ochakovski GA, Seitz IP, Wilhelm B, Ueffing M, Biel M, Wissinger B, et al. ; RD-CURE Consortium (2017). AAV8 can induce innate and adaptive immune response in the primate eye. Mol. Ther 25, 2648–2660. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rigotti DJ, Kirov II, Djavadi B, Perry N, Babb JS, and Gonen O (2011). Longitudinal whole-brain N-acetylaspartate concentration in healthy adults. AJNRAm. J. Neuroradiol 32, 1011–1015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rockwell HE, McCurdy VJ, Eaton SC, Wilson DU, Johnson AK, Randle AN, Bradbury AM, Gray-Edwards HL, Baker HJ, Hudson JA, et al. (2015). AAV-mediated gene delivery in a feline model of Sandhoff disease corrects lysosomal storage in the central nervous system. ASN Neuro 7, 7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rosario AM, Cruz PE, Ceballos-Diaz C, Strickland MR, Siemienski Z, Pardo M, Schob KL, Li A, Aslanidi GV, Srivastava A, et al. (2016). Microglia-specific targeting by novel capsid-modified AAV6 vectors. Mol. Ther. Methods Clin. Dev 3, 16026. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rosenberg JB, Kaplitt MG, De BP, Chen A, Flagiello T, Salami C, Pey E, Zhao L, Ricart Arbona RJ, Monette S, et al. (2018). AAVrh.10-mediated APOE2 central nervous system gene therapy for APOE4-associated Alzheimer’s disease. Hum. Gene Ther. Clin. Dev 29, 24–47. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Russell S, Bennett J, Wellman JA, Chung DC, Yu ZF, Tillman A, Wittes J, Pappas J, Elci O, McCague S, et al. (2017). Efficacy and safety of voretigene neparvovec (AAV2-hRPE65v2) in patients with RPE65-mediated inherited retinal dystrophy: a randomised, controlled, open-label, phase 3 trial. Lancet 390, 849–860. [DOI] [PMC free article] [PubMed] [Google Scholar]
- San Sebastian W, Samaranch L, Heller G, Kells AP, Bringas J, Pivirotto P, Forsayeth J, and Bankiewicz KS (2013). Adeno-associated virus type 6 is retrogradely transported in the non-human primate brain. Gene Ther 20, 1178–1183. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sanchez-Pernaute R, Harvey-White J, Cunningham J, and Bankiewicz KS (2001). Functional effect of adeno-associated virus mediated genetransfer of aromatic L-amino acid decarboxylase into the striatum of 6-OHDA-lesioned rats. Mol. Ther 4, 324–330. [DOI] [PubMed] [Google Scholar]
- Saraiva J, Nobre RJ, and Pereira de Almeida L (2016). Gene therapy for the CNS using AAVs: the impact of systemic delivery by AAV9. J. Control. Release 241, 94–109. [DOI] [PubMed] [Google Scholar]
- Schnepp BC, Jensen RL, Chen CL, Johnson PR, and Clark KR (2005). Characterization of adeno-associated virus genomes isolated from human tissues. J. Virol 79, 14793–14803. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Senís E, Fatouros C, Große S, Wiedtke E, Niopek D, Mueller AK, Borner K, and Grimm D (2014). CRISPR/Cas9-mediated genome engineering: an adeno-associated viral (AAV) vector toolbox. Biotechnol. J 9, 1402–1412. [DOI] [PubMed] [Google Scholar]
- Shibata SB, Ranum PT, Moteki H, Pan B, Goodwin AT, Goodman SS, Abbas PJ, Holt JR, and Smith RJH (2016). RNA interference prevents autosomal-dominant hearing loss. Am. J. Hum. Genet 98, 1101–1113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sondhi D, Johnson L, Purpura K, Monette S, Souweidane MM, Kaplitt MG, Kosofsky B, Yohay K, Ballon D, Dyke J, et al. (2012). Long-term expression and safety of administration of AAVrh.10hCLN2 to the brain of rats and nonhuman primates for the treatment of late infantile neuronal ceroid lipofuscinosis. Hum. Gene Ther. Methods 23, 324–335. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Srivastava A, and Carter BJ (2017). AAV infection: protection from cancer. Hum. Gene Ther 28, 323–327. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stein-Streilein J (2013). Mechanisms of immune privilege in the posterior eye. Int. Rev. Immunol 32, 42–56. [DOI] [PubMed] [Google Scholar]
- Strittmatter WJ, and Roses AD (1996). Apolipoprotein E and Alzheimer’s disease. Annu. Rev. Neurosci 19, 53–77. [DOI] [PubMed] [Google Scholar]
- Su X, Kells AP, Huang EJ, Lee HS, Hadaczek P, Beyer J, Bringas J, Pivirotto P, Penticuff J, Eberling J, et al. (2009). Safety evaluation of AAV2-GDNF gene transfer into the dopaminergic nigrostriatal pathway in aged and parkinsonian rhesus monkeys. Hum. Gene Ther 20, 1627–1640. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Takahashi K, Igarashi T, Miyake K, Kobayashi M, Yaguchi C, lijima O, Yamazaki Y, Katakai Y, Miyake N, Kameya S, et al. (2017). Improved intravitreal AAV-mediated inner retinal gene transduction after surgical internal limiting membrane peeling in cynomolgus monkeys. Mol. Ther 25, 296–302. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tallan HH (1957). Studies on the distribution of N-acetyl-L-aspartic acid in brain. J. Biol. Chem 224, 41–45. [PubMed] [Google Scholar]
- Tao Y, Huang M, Shu Y, Ruprecht A, Wang H, Tang Y, Vandenberghe LH, Wang Q, Gao G, Kong WJ, and Chen ZY (2018). Delivery of adeno-associated virus vectors in adult mammalian inner-ear cell subtypes without auditory dysfunction. Hum. Gene Ther 29, 492–506. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tardieu M, Zerah M, Husson B, de Bournonville S, Deiva K, Adamsbaum C, Vincent F, Hocquemiller M, Broissand C, Furlan V, et al. (2014). Intracerebral administration of adeno-associated viral vector serotype rh.10 carrying human SGSH and SUMF1 cDNAs in children with mucopolysaccharidosis type IIIA disease: results of a phase I/II trial. Hum. Gene Ther 25, 506–516. [DOI] [PubMed] [Google Scholar]
- Taschenberger G, Tereshchenko J, and Kugler S (2017). A MicroRNA124 target sequence restores astrocyte specificity of gfaABC1 D-driven transgene expression in AAV-mediated gene transfer. Mol. Ther. Nucleic Acids 8, 13–25. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Towne C, Schneider BL, Kieran D, Redmond DE Jr., and Aebischer P (2010). Efficient transduction of non-human primate motor neurons after intramuscular delivery of recombinant AAV serotype 6. Gene Ther 17, 141–146. [DOI] [PubMed] [Google Scholar]
- Trapani I, and Auricchio A (2018). Seeing the light after 25 years of retinal gene therapy. Trends Mol. Med 24, 669–681. [DOI] [PubMed] [Google Scholar]
- Tse LV, Klinc KA, Madigan VJ, Castellanos Rivera RM, Wells LF, Havlik LP, Smith JK, Agbandje-McKenna M, and Asokan A (2017). Structure-guided evolution of antigenically distinct adeno-associated virus variants for immune evasion. Proc. Natl. Acad. Sci. USA 114, E4812–E4821. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vandamme C, Adjali O, and Mingozzi F (2017). Unraveling the complex story of immune responses to AAV vectors trial after trial. Hum. Gene Ther 28, 1061–1074. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vandenberghe LH, and Auricchio A (2012). Novel adeno-associated viral vectors for retinal gene therapy. Gene Ther 19, 162–168. [DOI] [PubMed] [Google Scholar]
- Vandenberghe LH, and Wilson JM (2007). AAV as an immunogen. Curr. Gene Ther 7, 325–333. [DOI] [PubMed] [Google Scholar]
- Vandenberghe LH, Wilson JM, and Gao G (2009). Tailoring the AAV vector capsid for gene therapy. Gene Ther 16, 311–319. [DOI] [PubMed] [Google Scholar]
- Vandenberghe LH, Bell P, Maguire AM, Cearley CN, Xiao R, Calcedo R, Wang L, Castle MJ, Maguire AC, Grant R, et al. (2011). Dosage thresholds for AAV2 and AAV8 photoreceptor gene therapy in monkey. Sci. Transl. Med 3, 88ra54. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Walia JS, Altaleb N, Bello A, Kruck C, LaFave MC, Varshney GK, Burgess SM, Chowdhury B, Hurlbut D, Hemming R, et al. (2015). Long-term correction of Sandhoff disease following intravenous delivery of rAAV9 to mouse neonates. Mol. Ther 23, 414–422. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang LJ, Lu YY, Muramatsu S, Ikeguchi K, Fujimoto K, Okada T, Mizukami H, Matsushita T, Hanazono Y, Kume A, et al. (2002). Neuroprotective effects of glial cell line-derived neurotrophic factor mediated by an adeno-associated virus vector in a transgenic animal model of amyotrophic lateral sclerosis. J. Neurosci 22, 6920–6928. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang J, Lozier J, Johnson G, Kirshner S, Verthelyi D, Pariser A, Shores E, and Rosenberg A (2008). Neutralizing antibodies to therapeutic enzymes: considerations for testing, prevention and treatment. Nat. Biotechnol 26, 901–908. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang DB, Dayton RD, Henning PP, Cain CD, Zhao LR, Schrott LM, Orchard EA, Knight DS, and Klein RL (2010). Expansive gene transfer in the rat CNS rapidly produces amyotrophic lateral sclerosis relevant sequelae when TDP-43 is overexpressed. Mol. Ther 18, 2064–2074. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang PR, Xu M, Toffanin S, Li Y, Llovet JM, and Russell DW (2012). Induction of hepatocellular carcinoma by in vivo gene targeting. Proc. Natl. Acad. Sci. USA 109, 11264–11269. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang J, Exline CM, DeClercq JJ, Llewellyn GN, Hayward SB, Li PW, Shivak DA, Surosky RT, Gregory PD, Holmes MC, and Cannon PM (2015). Homology-driven genome editing in hematopoietic stem and progenitor cells using ZFN mRNA and AAV6 donors. Nat. Biotechnol 33, 1256–1263. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang L, Xiao R, Andres-Mateos E, and Vandenberghe LH (2017). Single stranded adeno-associated virus achieves efficient gene transfer to anterior segment in the mouse eye. PLoS ONE 12, e0182473. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wassmer SJ, Carvalho LS, György B, Vandenberghe LH, and Maguire CA (2017). Exosome-associated AAV2 vector mediates robust gene delivery into the murine retina upon intravitreal injection. Sci. Rep 7, 45329. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Watakabe A, Ohtsuka M, Kinoshita M, Takaji M, Isa K, Mizukami H, Ozawa K, Isa T, and Yamamori T (2015). Comparative analyses of adeno-associated viral vector serotypes 1, 2, 5, 8 and 9 in marmoset, mouse and macaque cerebral cortex. Neurosci. Res 93, 144–157. [DOI] [PubMed] [Google Scholar]
- Weismann CM, Ferreira J, Keeler AM, Su Q, Qui L, Shaffer SA, Xu Z, Gao G, and Sena-Esteves M (2015). Systemic AAV9 gene transfer in adult GM1 gangliosidosis mice reduces lysosomal storage in CNS and extends lifespan. Hum. Mol. Genet 24, 4353–4364. [DOI] [PMC free article] [PubMed] [Google Scholar]
- West HL, Rebeck GW, and Hyman BT (1994). Frequency of the apolipo-protein E epsilon 2 allele is diminished in sporadic Alzheimer disease. Neurosci. Lett 175, 46–48. [DOI] [PubMed] [Google Scholar]
- Whitehouse PJ, Price DL, Struble RG, Clark AW, Coyle JT, and Delon MR (1982). Alzheimer’s disease and senile dementia: loss of neurons in the basal forebrain. Science 215, 1237–1239. [DOI] [PubMed] [Google Scholar]
- Williams CL, Uytingco CR, Green WW, McIntyre JC, Ukhanov K, Zimmerman AD, Shively DT, Zhang L, Nishimura DY, Sheffield VC, and Martens JR (2017). Gene therapeutic reversal of peripheral olfactory impairment in Bardet-Biedl syndrome. Mol. Ther 25, 904–916. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wolf DA, Lenander AW, Nan Z, Belur LR, Whitley CB, Gupta P, Low WC, and McIvor RS (2011). Direct gene transfer to the CNS prevents emergence of neurologic disease in a murine model of mucopolysaccharidosis type I. Neurobiol. Dis 43, 123–133. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wolf DA, Hanson LR, Aronovich EL, Nan Z, Low WC, Frey WH 2nd, and McIvor RS (2012). Lysosomal enzyme can bypass the blood-brain barrier and reach the CNS following intranasal administration. Mol. Genet. Metab 106, 131–134. [DOI] [PubMed] [Google Scholar]
- Worgall S, Sondhi D, Hackett NR, Kosofsky B, Kekatpure MV, Neyzi N, Dyke JP, Ballon D, Heier L, Greenwald BM, et al. (2008). Treatment of late infantile neuronal ceroid lipofuscinosis by CNS administration of a serotype 2 adeno-associated virus expressing CLN2 cDNA. Hum. Gene Ther 19, 463–474. [DOI] [PubMed] [Google Scholar]
- Wrobel C, Dieter A, Huet A, Keppeler D, Duque-Afonso CJ, Vogl C, Hoch G, Jeschke M, and Moser T (2018). Optogenetic stimulation of cochlear neurons activates the auditory pathway and restores auditory-driven behavior in deaf adult gerbils. Sci. Transl. Med 10, 10. [DOI] [PubMed] [Google Scholar]
- Wu P, Phillips MI, Bui J, and Terwilliger EF (1998). Adeno-associated virus vector-mediated transgene integration into neurons and other nondividing cell targets. J. Virol 72, 5919–5926. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu Z, Yang H, and Colosi P (2010). Effect of genome size on AAV vector packaging. Mol. Ther 18, 80–86. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yamashita T, Chai HL, Teramoto S, Tsuji S, Shimazaki K, Muramatsu S, and Kwak S (2013). Rescue of amyotrophic lateral sclerosis phenotype in a mouse model by intravenous AAV9-ADAR2 delivery to motor neurons. EMBO Mol. Med 5, 1710–1719. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yang B, Li S, Wang H, Guo Y, Gessler DJ, Cao C, Su Q, Kramer J, Zhong L, Ahmed SS, et al. (2014). Global CNS transduction of adult mice by intravenously delivered rAAVrh.8 and rAAVrh.10 and nonhuman primates by rAAVrh.10. Mol. Ther 22, 1299–1309. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yin H, Kanasty RL, Eltoukhy AA, Vegas AJ, Dorkin JR, and Anderson DG (2014). Non-viral vectors for gene-based therapy. Nat. Rev. Genet 15, 541–555. [DOI] [PubMed] [Google Scholar]
- Yoshimura H, Shibata SB, Ranum PT, and Smith RJH (2018). Enhanced viral-mediated cochlear gene delivery in adult mice by combining canal fenestration with round window membrane inoculation. Sci. Rep 8, 2980. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zerah M, Piguet F, Colle MA, Raoul S, Deschamps JY, Deniaud J, Gautier B, Toulgoat F, Bieche I, Laurendeau I, et al. (2015). Intracerebral gene therapy using AAVrh.10-hARSA recombinant vector to treat patients with early-onset forms of metachromatic leukodystrophy: preclinical feasibility and safety assessments in nonhuman primates. Hum. Gene Ther. Clin. Dev 26, 113–124. [DOI] [PubMed] [Google Scholar]
- Zhang H, Yang B, Mu X, Ahmed SS, Su Q, He R, Wang H, Mueller C, Sena-Esteves M, Brown R, et al. (2011). Several rAAV vectors efficiently cross the blood-brain barrier and transduce neurons and astrocytes in the neonatal mouse central nervous system. Mol. Ther 19, 1440–1448. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhong L, Malani N, Li M, Brady T, Xie J, Bell P, Li S, Jones H, Wilson JM, Flotte TR, et al. (2013). Recombinant adeno-associated virus integration sites in murine liver after ornithine transcarbamylase gene correction. Hum. Gene Ther 24, 520–525. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zingg B, Chou XL, Zhang ZG, Mesik L, Liang F, Tao HW, and Zhang LI (2017). AAV-mediated anterograde transsynaptic tagging: mapping corticocollicular input-defined neural pathways for defense behaviors. Neuron 93, 33–47. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zinn E, Pacouret S, Khaychuk V, Turunen HT, Carvalho LS, AndresMateos E, Shah S, Shelke R, Maurer AC, Plovie E, et al. (2015). In silico reconstruction of the viral evolutionary lineage yields a potent gene therapy vector. Cell Rep 12, 1056–1068. [DOI] [PMC free article] [PubMed] [Google Scholar]