Next Article in Journal
Micrandilactone C, a Nortriterpenoid Isolated from Roots of Schisandra chinensis, Ameliorates Huntington’s Disease by Inhibiting Microglial STAT3 Pathways
Next Article in Special Issue
Transcriptome-Powered Pluripotent Stem Cell Differentiation for Regenerative Medicine
Previous Article in Journal
Context-Dependent Role of Glucocorticoid Receptor Alpha and Beta in Breast Cancer Cell Behaviour
Previous Article in Special Issue
Olfactory Regeneration with Nasally Administered Murine Adipose-Derived Stem Cells in Olfactory Epithelium Damaged Mice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Various AAV Serotypes and Their Applications in Gene Therapy: An Overview

by
Shaza S. Issa
1,
Alisa A. Shaimardanova
2,
Valeriya V. Solovyeva
2 and
Albert A. Rizvanov
2,*
1
Department of Genetics and Biotechnology, St. Petersburg State University, 199034 St. Petersburg, Russia
2
Institute of Fundamental Medicine and Biology, Kazan Federal University, 420008 Kazan, Russia
*
Author to whom correspondence should be addressed.
Cells 2023, 12(5), 785; https://doi.org/10.3390/cells12050785
Submission received: 29 December 2022 / Revised: 22 February 2023 / Accepted: 26 February 2023 / Published: 1 March 2023
(This article belongs to the Special Issue Gene and Cell Therapy in Regenerative Medicine)

Abstract

:
Despite scientific discoveries in the field of gene and cell therapy, some diseases still have no effective treatment. Advances in genetic engineering methods have enabled the development of effective gene therapy methods for various diseases based on adeno-associated viruses (AAVs). Today, many AAV-based gene therapy medications are being investigated in preclinical and clinical trials, and new ones are appearing on the market. In this article, we present a review of AAV discovery, properties, different serotypes, and tropism, and a following detailed explanation of their uses in gene therapy for disease of different organs and systems.

1. Introduction

Thanks to the rapid progress in biomedical technologies, large developments of novel high-tech treatment approaches have become possible today for several human diseases. However, despite scientific discoveries in the field of gene and cell therapy, some diseases still have no effective treatment. Advances in genetic engineering methods have enabled the development of effective gene therapy methods for various diseases based on adeno-associated viruses (AAVs). AAVs are effectively used to treat a number of genetic and acquired human diseases, more specifically, monogenic hereditary ones, i.e., diseases caused by mutations in one gene. Today, many AAV-based gene therapy medications are being investigated in preclinical and clinical trials, and new ones are appearing on the market for the treatment of numerous human diseases, including those that were previously considered incurable.
To date, there are only a few approved AAV-based gene therapy medications. For example, the US Food and Drug Administration (FDA) recently approved “Hemgenix”, a drug for the treatment of hemophilia type B, which is caused by congenital deficiency of factor IX. The drug is an AAV serotype 5 (AAV5), carrying the gene of deficient factor IX. As for hemophilia type A, Roctavian has also recently been approved by the FDA as an AAV5-based gene therapy vector, carrying the deficient factor VIII gene, driven by a liver-selective promoter [1]. Another breakthrough gene therapy medication is “Zolgensma”, which was intended for spinal muscular atrophy (SMA) treatment, a disease characterized by degeneration of the anterior horns motor neurons in the spinal cord resulting in a loss of their functions. Zolgensma is an AAV-serotype-9 (AAV9)-based vector encoding complementary DNA (cDNA) of the survival motor neuron 1 gene (survival motor neuron, SMN1). The drug was approved by the FDA in May 2019 for the treatment of type 1 SMA. Following intravenous administration, AAV9 crosses the blood–brain barrier (BBB) and provides targeted delivery of the SMN gene to neurons. It has been shown that intravenous administration of Zolgensma leads to an improvement in patients’ motor skills, along with a decrease in the clinical manifestations of SMA [2]. Another FDA-approved AAV-based gene therapy medication is Luxturna, which is intended for the treatment of retinal dystrophy caused by biallelic mutations in the RPE65 gene (an enzyme of retinal cells involved in light-sensitive pigment regeneration). Drug delivery to the retinal cells leads to synthesis restoration of normal RPE65 protein, which is necessary for the regeneration of the photosensitive pigment [3]. The very first AAV-based gene therapy medication to be registered is Glybera, which was approved by the European Commission for the treatment of a rare hereditary disease, lipoprotein lipase deficiency [4].
Nevertheless, despite the successful development of AAV-based gene therapy medications, many preclinical and clinical studies still get suspended or terminated, without resulting in a drug registration, and therefore, many diseases, especially the rare ones, remain untreated. One explanation for that could be the lack of knowledge in gene therapeutics, in particular, features of AAV vectors. To tackle this issue, further research is needed to achieve better optimization and higher effectiveness of AAV-based gene therapy approaches. This review discusses various AAV serotypes, some features of infection, tissue and organ tropism, efficacy in many human diseases, etc.

2. Brief History of AAVs’ Discovery

AAVs are small non-enveloped DNA viruses belonging to the Parvoviridae family, that were first isolated in 1965, as a contaminant in preparations of a simian adenovirus (Ad) [5,6]. The viruses were found incompetent to productively infect cells without a co-infection by a helper virus, usually an Ad or any type of herpesviruses, and thereby they were named as “adeno-associated”, and classified into the Dependovirus genus [7,8]. After being regarded as defective viruses for a long time due to their codependency, later studies on AAVs disproved this theory and showed that they rather launch a latent infection in the host cell, that could convert to productive infection under stress [8,9]. Although AAVs have a high seroprevalence in humans (it has been estimated that 50% to 96% of the human population is seropositive for the second serotype of AAV (AAV2) depending on age and ethnic group) [10,11,12], they, however, were not linked to any disease neither in humans, nor in any other species [8]. Different AAVs have not only been detected in primates isolates [13], but also from avian [14,15,16], caprine [17,18], bovine [19,20], and equine stocks [21].

3. Properties, Structure, and Genome Organization of AAVs

Aside from AAV5 being the most divergent, all AAVs share a similar structure and properties [22]. AAVs are easy to manipulate, as their particles can maintain biological stability in extreme conditions of pH and temperature [23,24]. They share a genome of approximately 4.7 kb single-stranded DNA packed into an icosahedral, non-enveloped capsid with a diameter of 20–25 nm [6,25]. The AAV genome consists mainly of two viral genes: rep (replication) and cap (capsid), flanked by inverted terminal repeats (ITRs) [12,25]. As the ITRs have a palindromic nucleotide sequence, they create characteristic T-shaped hairpin structures, providing essential structural elements for viral genome replication and packaging [26]. ITRs also play a regulatory role in viral gene expression and host genome integration [10,27]. The open reading frame (ORF) of rep encodes several nonstructural proteins that are required for gene regulation, replication, transcription, and encapsidation [12,28,29], while the ORF of cap encodes three structural proteins: virion protein 1 (VP1), VP2, and VP3, with a molar ratio of 1:1:10 in AAV particles [10,30,31]. Distinct tissue tropism of different AAV serotypes results from variations in the processing of this cap ORF, leading to variant immune and transduction profiles [12,32].

4. AAV Serotypes and Tropism

Depending on their serotype, AAVs can have specific tropism for specific organs and tissues of the body. There are different AAV serotypes that vary in many aspects. Next, each serotype will be separately discussed. Table 1 summarizes the characteristics and properties of AAV serotypes, and Figure 1 demonstrates their variant tropisms.

4.1. AAV1

The exact origin of the first serotype of AAV (AAV1) is still unknown, as it was not initially isolated from tissues, but as a contaminant of Ad stocks [33,34], and its antibodies were found both in humans and non-human primates (NHPs) [35]. This serotype uses sialic acid as its primary cellular surface receptor [36], and AAV receptor (AAVR) as a coreceptor [37]. According to Rabinowitz J.E. et al., AAV1 does not bind heparin as it lacks R585 and R588, the amino acid residues required for such binding, and thereby it cannot be purified using heparin [33,38]. Zolotukhin S. et al., developed a protocol for AAV1 chromatographic purification by iodixanol gradient centrifugation and anion-exchange chromatography [39]. It can also be purified using mucin columns, as it can bind the sialic acid residues in mucin [40]. Moreover, recombinant AAV1 (rAAV1) was not found to contain any detectable post-translational modifications (PTMs), according to a systematic analysis conducted on ten AAV serotypes by Mary B. et al. [41], and it was the first viral vector to be approved for use in gene therapy [42]. In 1999, Xiao W. et al., conducted a study to investigate viral vectors for gene therapy, as a result of which, AAV1 was found to be the most efficient serotype for skeletal muscles’ transduction [35]. Many studies have since confirmed AAV1 high tropism to skeletal muscles of murine, canine, and NHP origins, compared to other serotypes [43,44,45,46]. AAV1 was also found to achieve an efficient transduction of neurons and glial and ependymal cells in the murine brain [47]. Moreover, it was found to be able to effectively transduce the heart [48,49], endothelial and vascular smooth muscles [50], and retina [51].

4.2. AAV2

AAV2 is considered the most studied serotype among all AAVs [45]. It was first discovered in 1965 as a contaminant of simian Ad preparations [52]. Later, in 1998, its primary cellular receptor, heparan sulfate proteoglycan (HSPG), was identified by Summerford C. and Samulski R.J. [53], and the amino acid residues providing its affinity to HSPG were suggested, afterwards, as R585 and R588 [54]. Accordingly, rAAV2 can be purified using heparin column affinity chromatography [55]. Nevertheless, the binding of AAV2 to its primary receptor was found to be insufficient for cell entry, so several coreceptors were later identified for it [52], including the human fibroblast growth factor receptor 1 (FGFR1) [56], αVβ5 [57] and α5β1 integrins [58], hepatocyte growth factor receptor (HGFR) [59], laminin receptor (LR) [60], and CD9 [61]. The capsid of rAAV2 is reported to acquire multiple PTMs, including ubiquitination, phosphorylation, SUMOylation, and multiple-site-glycosylation [41]. As the most studied serotype, AAV2, in fact, has shown various tropisms for several tissues in NHP, murine, canine, avian, and human cell types, including renal tissue [62,63,64], hepatocytes [65,66], retina [67,68,69], non-mitotic cells of the central nervous system (CNS) [70,71], and skeletal muscles [45,71]. Nonetheless, broader tissue tropisms of AAV2 were enabled by the further innovation of mosaicism and cross-packaging (or cross-typing) of AAVs (explained in detail later in this article), where the viral genome of one serotype could be packaged into a capsid of another type, providing a wider transduction spectrum [38,72,73]

4.3. AAV3

The third serotype of AAV (AAV3) was originally isolated from humans [34,45]. Similar to AAV2, this serotype uses HSPG, FGFR1, and LR receptors [60,74], along with the human HGFR (hHGFR) receptor [60]. Iodixanol gradient ultracentrifugation along with ion exchange chromatography have been used for AAV3 purification [75]. PTMs of rAAV3 capsid include acetylation, phosphorylation, and glycosylation [41]. Due to its inadequate transduction efficiency in vitro and in murine cell lines, AAV3 was mostly overlooked as a choice for gene therapy [76]. However, as it has been later found to use hHGFR as a coreceptor, it showed extremely efficient transduction of human liver cancer cells as well as human and NHP hepatocytes [76,77]. Since this selective tropism of AAV3 has been discovered, various studies have aimed at optimizing strategies to generate rAAV3 vectors with a higher transduction efficiency [76,78,79,80,81]. The developed strategies suggested different approaches, mainly, capsid modification of AAV3 vectors [76,80,81] and modification of hHGFR expression levels, along with the tyrosine kinase activity associated with it [78]. AAV3 was also found to have specific tropism to cochlear inner hair cells, showing high in vivo transduction efficiency in a murine model [82].

4.4. AAV4

The fourth serotype of AAV (AAV4) is considered one of the most antigenically distinct serotypes [83]. It has reportedly originated in NHPs [84], mainly in green African monkeys [85], as antibodies to its viral particles have been detected in their sera [86,87]. A study of the AAV4 structure showed that its capsid surface topology shares a significant similarity with that of human parvovirus B19 and Aleutian mink disease virus [83]. AAV4 uses a-2,3-O-linked sialic acid for cell binding and infection [88]; accordingly, mucin columns can be used for AAV4 purification, based on its ability to bind sialic acid residues in mucin [33,40]. In addition, as this serotype lacks heparin-binding activity, it cannot be purified using heparin column affinity chromatography like AAV2, however, ion-exchange chromatography procedures have been developed and proven a high purification efficiency [89]. The only reported PTM of rAAV4 is ubiquitination of its capsid proteins [41]. AAV4 is suggested to be able to transduce human/NHP cells, as well as cells of murine and canine origins [45,87]. The specific tropism of AAV4 results in a transduction efficacy of specific cell types in the mammalian central nervous system (CNS), mainly the ependymal cells [90]. Moreover, following subretinal delivery, AAV4 has showed stable transduction of retinal pigmented epithelium (RPE) cells of rodent, canine, and nonhuman primate models, a distinctive feature enabled by the specificity of its capsid [91]. In a murine model, AAV4 has also shown a significant transduction rate of kidney, lung, and heart cells [92].

4.5. AAV5

As it was first isolated in 1983 from male genital lesions, AAV5 became the only AAV serotype to be isolated directly from a human tissue [93]. This serotype is considered the most genetically divergent of all AAVs [6,94,95,96], with a variety of unique characteristics, such as the distinct size and function of its ITR regions [22], utilizing herpes simplex virus (HSV) as its helper virus for human infections [94], and using an atypical endocytic route as a pathway for viral entry [33]. Another distinctive feature of AAV5 is its ability to transduce cells that cannot be transduced with AAV2, an exclusive advantage for gene therapy uses [97,98,99]. AAV5 was also found to use sialic acid as its primary receptor [88,100,101], along with platelet-derived growth factor receptors (PDGFR) α and β as coreceptors [102,103]. Similar to AAV1 and AAV4, mucin columns can be used for AAV5 purification [104,105], and ion-exchange chromatography procedures have also been developed for that aim [89]. Capsid proteins of rAAV5 are reported to undergo multiple PTMs, including ubiquitination, phosphorylation, SUMOylation, and glycosylation [41]. AAV5 proved to have a significant transduction efficiency for murine retinal cells [38,51], mainly for photoreceptor cells [106]. In addition, AAV5 tropism has been investigated in the murine brain, and proved, as a result, a transduction competence for multiple neural cell types, including Purkinje cells, stellate, basket, and Golgi neurons, and it was able to reach to the inferior colliculus and ventricular epithelium [107,108]. AAV5 is also known for its efficient transduction of murine airway epithelia by apical infection [98,109,110,111], vascular endothelial cells, and smooth muscles [50]. It is also reported to have tropism for liver cells in mice [92,112].

4.6. AAV6

The classification of the sixth serotype of AAV (AAV6) is still a matter of controversy, as it presents high genomic similarity with both AAV1 and AAV2 serotypes, however, it has still been assigned its own serotype numbering [33,35]. AAV6 has a serological profile almost identical to that of AAV1, and shares its sequence of coding region with a homology percentage of 99%, along with multiple regions identical to those of AAV2 [113]. Accordingly, it was suggested to be a naturally occurring hybrid resulting from homologous RECOMation between AAV1 and AAV2 [35,45,113,114]. AAV6 was first isolated from a human Ad preparation [45,113,115], and similar to AAV1, was found to bind sialylated proteoglycans, mainly α2,3-/α2,6-linked sialic acid, as its primary receptor, as well as binding heparan sulfate [36,116,117]. As for its coreceptor, it binds epidermal growth factor receptor (EGFR) [118]. The only reported PTM of rAAV6 is acetylation of its capsid proteins [41]. Similar to the previously described serotypes, AAV6 can be purified by either heparin or mucin column affinity chromatography, as it can bind both [40]. AAV6 is reported to have tropism for a variety of tissues, including airway epithelia of murine and canine models [119,120], murine liver cells [92,121], and skeletal muscles of murine and canine models, with a transduction efficiency even higher than that of AAV2 [92,122,123,124], cardiomyocytes in murine [92,125], porcine [126], canine [127], and in sheep [128] models.

4.7. AAV7

The seventh serotype of AAV (AAV7) was first isolated in 2002 from NHP tissues, specifically, from Rhesus macaque monkeys [35,129]. Its mechanisms of cell binding and cell entry are still unknown [52,130], but it is established that this serotype does not bind heparin, or any other glycan in general [131]. Capsid proteins of rAAV7 undergo multiple PTMs, including glycosylation, primarily, along with phosphorylation, SUMOylation, and acetylation [41]. A study by Calcedo R et al., investigated the epidemiology of AAV-neutralizing antibodies in the worldwide population, and found that the seroprevalence of AAV7 antibodies is relatively low in humans, an advantage of this serotype to be used in clinical applications [132]. Viral vectors based on AAV7 proved a high efficiency of transduction for skeletal muscle cells in murine models, similar to that achieved by AAV1, and higher than AAV2 [133]. This serotype also proved a strong tropism to hepatocytes in murine [92] and human [92] tissues. In CNS of NHPs, AAV7 viral vectors were found to achieve a robust transduction mainly in cortical and spinal tissues [134]. Moreover, AAV7-based viral vectors can, reportedly, achieve a significantly high transduction rate of murine neurons [135] and photoreceptor cells in the retina both in vivo and ex vivo [136]. AAV7 vectors also appear to have a limited tropism to vascular endothelial cells, that could be relatively enhanced through proteasome inhibition [130], and an in vivo transduction preference for epicardium cells in murine cardiac tissue [49].

4.8. AAV8

Similar to AAV7, the eighth serotype of AAV (AAV8) was first isolated in 2002 from Rhesus macaque monkeys [35,129]. As a primary receptor, AAV8 binds LR, the same receptor used by AAV2 and AAV3 [45,60]. Various procedures for rapid and scalable purification of AAV8 have been developed since its discovery, including, for example, dual-ion-exchange chromatography [137,138], or iodixanol gradient centrifugation [139,140]. Phosphorylation, glycosylation, and acetylation are the three PTMs reported for rAAV8 capsid proteins [41]. AAV8 is best known for its strong tropism to liver cells, and accordingly, its transduction efficiency of hepatocytes, which is far stronger and faster than those of all other AAV serotypes in different models, including murine, canine, and NHP [92,112,141,142,143,144,145,146,147]. Following systemic delivery in murine models, AAV8 was proven to be the most efficient serotype for transduction of both skeletal and cardiac muscles, owing to its ability to cross the blood vessel barrier, a feature that both AAV1 and AAV6 lack, limiting their efficiency of muscle transduction to local delivery only [148]. AAV8 could also achieve a successful in vivo transduction of murine pancreatic cells, following localized delivery [149,150]. Moreover, a high-rate transduction of murine renal cells could be reached by local direct delivery of AAV8 viral vectors into the kidney tissue [151]. AAV8 was also found to achieve an efficient transduction of different cells in the murine retina, including amacrine, Müller, and putative bipolar cells, along with some horizontal cells and cells in the ganglion cell layer (GCL) [152,153,154]. AAV8 transduction efficiency appears to be susceptible to proteasome levels in some tissues, and accordingly, could be increased using proteasome inhibitors [130].

4.9. AAV9

The ninth serotype of AAV (AAV9) was first identified in a human isolate in 2004, and was named a new serotype as it had a serological profile distinct from the previously known AAVs, however, it was suggested to be closely related to clades containing AAV7 and AAV8 [129,155]. As a primary receptor, AAV9 uses terminal N-linked galactose [156,157], and it is also suggested to bind a putative integrin, along with LR as coreceptors [60,156]. Scalable simple purification protocols have been developed for AAV9 purification, including ion-exchange chromatography [24] and sucrose gradient centrifugation [158]. Capsid of rAAV9 has one of the highest totals of PTMs, including multiple ubiquitination, phosphorylation, SUMOylation, and glycosylation modifications, along with acetylation [41]. In most tissues, AAV9 seems to achieve cell transduction with efficiency superior to other AAVs [33,129]. For example, in a study aimed to investigate AAV1–9 distribution following systemic delivery in a murine model, AAV9 has shown rapid-onset, the best genome distribution, and the highest protein levels, in comparison with all other AAVs [92]. Moreover, in the CNS of murine, NHP, and feline models, it has a unique feature compared to other serotypes, as it can cross the BBB and transduce not only neuronal but also non-neuronal cells, including astrocytes, that cannot usually be transduced by other AAVs [27,159,160,161,162], also showing tropism to photoreceptor cells in the retina [163]. Viral vectors based on AAV9 have also proven to be more efficient than those of AAV1 and AAV8 (in some cases 5–10-fold higher than AAV8), for murine, NHP, and porcine cardiac muscle transduction [164,165,166,167,168,169], enabled with another distinctive feature of AAV9—its ability to traverse the physical barrier of vascular system endothelia [168]. Another example of AAV9 superiority over other AAVs was presented in a study by Inagaki K et al., where the serotype achieved robust transduction of murine hepatocytes, skeletal muscles, and pancreatic cells [81,169]. AAV9-based viral vectors also appear to be tropic to murine photoreceptor cells [170], renal tubular epithelium cells [171,172,173], Leydig cells in the testicular interstitial tissue [174], and alveolar and nasal epithelia [175,176].

4.10. AAV10 and AAV11

The ninth and tenth serotypes of AAV (AAV10 and AAV11) were first found and described in 2004 in NHP isolates, namely from cynomolgus monkeys, with capsid proteins of great resemblance to AAV8 and AAV4, respectively [34,177], resulting in serological cross-reactivity with those two serotypes [178]. However, antisera against AAV10 and AAV11 were not found to have any cross-reactivity against those of AAV2, which recommended them as good viral vector candidates for gene therapy in individuals having antibodies against the latter [33]. It remains unknown what cellular receptors and coreceptors AAV10 and AAV11 use for cell binding and entry [37,115]; therefore, procedures describing their purification protocols are generally based on iodixanol gradient centrifugation [179]. Similar to AAV9, capsid of rAAV10 has one of the highest totals of PTMs, including mainly multiple glycosylation and phosphorylation modifications, along with ubiquitination, SUMOylation, and acetylation [41]. A study investigating biological distribution of both AAV10 and AAV11 in monkeys suggested them to be tropic to NHP intestinal cells, hepatocytes, lymph nodes, and less frequently, to renal cells and adrenal glands [180]. AAV10 was also suggested to have tropism to murine small intestine and colon cells [181]. Compared to AAV8 and AAV9, AAV10 appears to have the largest tropism range to murine retinal cells, as it can reportedly transduce a variety of cell types, including RPE, cells in the ganglion cell layer, several cell types in the inner nuclear layer, photoreceptors, and a highly efficient transduction of horizontal cells [152]. Following intravenous delivery, AAV10 was found to target murine liver and lung cells [182], however, upon localized delivery, it transduced murine renal and pancreatic cells [183]. As for AAV11, it was found to have mild tropism to NHP CNS, mainly to the cerebrum and spinal cord [180]. A recent study of neural gene therapy using rAAV11 found that murine projection neurons and astrocytes could also be targeted using this serotype [184].

4.11. AAV12

The twelfth serotype of AAV (AAV12) was first isolated from a simian Ad stock, and then characterized as a novel serotype, as it exhibited distinctive biological and serological properties [185]. Although it was proven that AAV12 does not use heparan sulfate proteoglycans or sialic acids for attachment and cell entry, it remains unknown how exactly it binds target cells [177,185]. However, according to a study investigating components of a potential receptor complex for AAV12, mannose and mannosamine were suggested as components of such complex, as they inhibited AAV12 cell transduction [186]. Moreover, being resistant to neutralization by human antibodies, AAV12 represents a good candidate for human gene therapy applications [185,187]. AVB Sepharose affinity chromatography has been used for purification of AAV12 [188]. In murine models, it has shown tropism to salivary glands and muscles [185]. It has also shown strong, localized in vivo tropism to murine nasal epithelia following intranasal administration [187].

4.12. AAV13

The thirteenth serotype of AAV (AAV13) is another simian Ad that appears to bind HSPG, although its primary cell receptor remains unknown [189,190,191]. It was also found to share structural similarity with AAV2 and AAV3, making it the closest related AAV to those two serotypes, with a capsid conserving all AAV capsids’ structural features [188], but there are only limited data on this serotype tropism and transduction efficiency [191].

4.13. Novel Hybrid AAV Vectors

In addition to the natural AAV serotypes described above, novel AAV vectors have been developed during the last two decades, and are still being developed [192]. Using different engineering strategies, novel hybrid vectors have been generated in order to enhance their transduction, modulate their immunogenicity, or limit their tropism to specific cells [193,194,195,196,197]. There has been different types of such engineered novel vectors, including mosaic, chimeric, and combinatorial vector libraries [33]. Mosaic vectors have multiple subunits of various serotypes in their capsid, chosen according to their properties of receptor binding and intracellular trafficking [33,198]. In chimeric virions, the capsid usually has modified protein generated by domain swapping and DNA shuffling strategies to alter specific amino acids [33,194]. Combinatorial vector libraries also use DNA shuffling and error-prone PCR methods to generate AAV libraries of novel serotypes with mixed genomes [199,200].

5. AAV as Viral Vectors for Gene Therapy Applications

A variety of AAV features have made it an appealing viral vector candidate to be used in gene therapy applications. To begin with, AAVs are non-pathogenic viruses, since no diseases have been linked to them [201]. AAV vectors have also proven their ability to provide stable integration into the target genome [202,203,204,205], and such feature is highly needed for gene therapeutic uses considering the high level of expression that could be maintained as a result [206], without affecting the target cell functions in the long term [207,208,209]. Moreover, having only ITRs, AAV vectors cannot interfere with the inserted gene regulation either [201]. More specifically, a unique feature of AAV vectors, contrary to all currently available gene-editing platforms, is utilizing the homologous recombination pathway, which does not involve exogenous nucleases, providing, therefore, a highly accurate editing process that preserves genome integrity without adding to the mutational burden of the target site [210]. Thanks to such unique gene-editing properties, AAV vectors are currently the leading platform for in vivo gene therapy delivery [210,211]. Another attractive feature is the very broad range of cells, tissues, and hosts that AAV can efficiently transduce in vivo and in vitro [87,201], including dividing and non-dividing cells [212,213] in humans [99], and NHP [214,215], murine [216,217,218], canine [219,220,221], feline [222,223], and a variety of other models. In addition to the genetic properties, AAVs’ distinct physical properties represent more reasons why they should be used for gene therapy, such as easy manipulation, resistance to pH changes, heat changes, and detergents [87]. Next, we present various studies that investigated AAV vectors for a wide range of gene therapy applications (reviewed in Table 2).

5.1. AAV Viral Vectors for Gene Therapy of the CNS

AAV-mediated CNS gene therapy was first believed to predominantly target neurons, with lesser chance to affect other cells and tissues in the CNS [224], a fact that was later proven to be inaccurate in many studies [134,160,214,225]. For a variety of neurodevelopmental and neurodegenerative diseases, AAV-mediated gene therapy has been tested using different administration routes, including intraparenchymal, intrathecal, intracerebroventricular, and intracisternal injection, many of which have shown promising results [226].
Parkinson’s disease (PD) has been one of the most studied neurological disorders as a target for AAV-mediated gene therapy [224,227,228]. In a phase I clinical trial, Kaplitt et al., investigated an AAV-mediated gene therapy approach for advanced PD patients, where serotype 2 was used as a vector for unilateral subthalamic delivery of the glutamic acid decarboxylase (GAD) gene [229]. Providing a significant improvement in motor function scores up to 12 months after surgery, the approach was found to be safe and well-tolerated in patients. A similar double-blind, controlled, randomized clinical trial conducted later by LeWitt et al., investigated the same AAV2-GAD vector for bilateral subthalamic delivery in advanced PD patients, and showed similar results of safety and improvement of motor function [230]. Bartus et al., also tested the bilateral stereotactic delivery of AAV2-neurturin in PD patients of an open-label clinical trial, the initial obtained data of which supported the feasibility, safety, and good tolerance of the approach as a potential treatment for PD [231]. However, bilateral intra-striatal infusion of an AAV2 vector containing the aromatic L-amino acid decarboxylase (AADC) gene in moderately advanced PD patients led to an improvement in PD rating scales that was associated with a risk of intracranial hemorrhage in patients, along with headaches [232]. As for sustainable transgene expression, a clinical trial conducted by Mittermeyer G. et al., investigated the potentials of rAAV2 carrying the aromatic L-amino acid decarboxylase gene (rAAV2-AADC) [233]. In the trial, 10 patients with moderately advanced PD received bilateral infusions of recombinant vector into the putamin, showing good tolerance and a stable expression of transgene that lasted for the following 4 years, although higher vector doses were suggested for further studies. In children with AADC deficiency, the same delivery route for the same vector (rAAV2-AADC) was also assessed in an open-label phase I/II clinical trial [234]. The therapy was well-tolerated in general, providing evidence for potential improvement of motor function.
Similarly, spinal muscular atrophy (SMA) has always represented an attractive candidate for gene therapy, being caused by a single gene defect that affects the survival motor neuron (SMN) protein [235]. Therefore, SMA has been suggested as a target for AAV-mediated therapeutic approaches, mainly using AAV9 [161]. An open-label, phase I clinical trial was designed by AveXis, Inc. in 2014 to assess the safety and efficacy of AA9-SMN as a treatment for SMA [236]. Following single-dose intravenous administration of the vector, a significant improvement of the motor function in all 15 patients was observed, reflected by their ability to perform different activities, such as unassisted sitting and walking, oral feeding, and speaking, with no reported motor function regression at the two-year follow-up [236]. A long-term safety assessment, however, was recommended.
Lysosomal storage diseases (LSDs), another group of neurodegenerative diseases, have also been extensively targeted by AAV-mediated gene therapy [224]. One example of which is mucopolysaccharidosis type VII (MPS VII or Sly disease), a disease that results from genetic deficiency of beta-glucuronidase (GUSβ) [237]. In the murine model, AAV-mediated gene therapy has been shown to provide stable expression of the deficient enzyme upon single administration, that was adequate for phenotype correction in the liver and most of the neuraxis [238,239]. As for neonatal murine model, intravenous delivery of rAAV-GUSβ proved to yield therapeutic levels of the deficient enzyme in multiple organs, including ones of the CNS, with the gene expression not being affected by rapid growth and differentiation of tissues [240]. The canine model of MPS VII has also shown promising therapy results following rAAV-GUSβ intrathecal injection of serotypes 9 and the NHP-derivate AAVrh10 [221], as high expression levels of GUSβ were detected in the CNS tissues, with the enzyme in brain tissue homogenates showing over 100% normal activity. Another type of MPS, MPS IIIA, has also been assessed for gene therapy in the canine model using AAV9 carrying the deficient enzyme’s gene [241]. Following intra-cerebrospinal fluid (CSF) delivery, the administered vector provided sustained and widely distributed transgene expression with no toxicity for a duration of seven years after therapy. Metachromatic leukodystrophy (MLD) is also one of the LSDs that results from arylsulfatase A (ARSA) enzyme deficiency and has been targeted by AAV-mediated therapeutic approaches [242]. In the murine model of MLD, as well as in NHPs, serotypes 1, 5, 9, and rh10 carrying the ARSA gene have provided supporting results of their therapeutic potentials as gene therapy viral vectors [243,244,245,246]. GM2 gangliosidoses is another example of lysosomal storage diseases that have been assessed for AAV-mediated gene therapy, including Tay-Sachs disease (TSD) and Sandhoff disease (SD) [247,248,249]. Preclinical data from many in vivo studies have provided evidence of AAVs’ efficacy for GM2 gangliosidoses in murine [250], ovine [251], and feline [252] models. Clinically, there has been trials applying AAV-mediated gene therapy for different LSDs throughout the last decades with variant outcomes [242,248,253]. In 2018, four children with asymptomatic or early-stage MLD underwent a phase I/II trial for gene therapy using AAVrh.10-hARSA [253]. Although there was a significant elevation in ARSA levels recorded in the CSF following intracerebral vector delivery, no clinical improvement was noticed compared to the control group. As for gangliosidosis, the first clinical trial for gene therapy of Tay-Sachs disease using an AAV vector was conducted by Taghian et al., in 2020 [248]. The treatment proved to be safe for the two treated children and resulted in a broad distribution of the transgene in CNS.
Canavan disease (CD) is another rare inherited leukodystrophy belonging to LSDs, that has been targeted by AAV-mediated gene therapy [254]. CD results from aspartoacylase (ASPA) enzyme deficiency, that leads to N-acetylaspartic acid (NAA) accumulation in the brain, causing white matter degeneration [255]. In a murine model, rAAV2-ASPA delivery into the striatum and thalamus of the brain resulted in an increased ASPA activity and, thereby, decreased NAA accumulation and white matter degeneration. However, areas remote from the injection site, such as the cerebellum, were not affected [255,256]. In 2002, Janson et al., suggested a clinical protocol using rAAV2-ASPA for gene therapy in CD patients [257], and currently there are several ongoing clinical trials testing different serotypes of rAAV (including 2, 9, and olig001), carrying the ASPA gene for CD gene therapy [258,259,260].
Krabbe disease, also known as globoid cell leukodystrophy (GLD), causes demyelination in a similar way. The diseases results from deficiency of the lysosomal enzyme galactocerebrosidase (GALC), leading to accumulation of its substrate, psychosine, which affects both the central and peripheral nervous systems [261]. During the last two decades, a large number of in vivo studies have been conducted to investigate different AAV serotypes as viral vectors for gene therapy of GLD in different animal models, including murine and canine. Intracerebral delivery of rAAV1-GALC has been tested in the murine model of GLD in an in vivo study conducted by Rafi et al. [262]. The approach resulted in sustained expression of the deficient enzyme, and thereby positively affected myelination status, and animals’ lifespan. However, both treated and untreated mice died with similar symptoms, suggesting that the used approach should be initiated prior to symptoms’ onset. rAAV2/5-GALC has also shown therapeutic efficacy in the murine model of GLD following intracranial delivery [263,264]. In the study conducted by Lin et al., in 2011, mice treated with rAAV2/5-GALC showed wide dispersion of the GALC transgene across the CNS, reaching areas remote from the injection site [263]. Moreover, rAAV2/5-GALC delivery resulted in a reduced loss of oligodendrocytes and Purkinje cells, along with a significant improvement of neuromotor function and a prolonged lifespan of treated mice. Karumuthil-Melethil et al., conducted another preclinical study using recombinant vectors AAV9, AAVrh10, and AAV-Olig001, carrying the GALC gene, for GLD gene therapy in the murine model, following lumbar intrathecal delivery [265]. All three serotypes provided wide distribution of the transgene across the CNS and liver, resulting in a significant improvement of myelination, and a prolonged lifespan, with AAV9 being the most effective when combined with bone marrow transplantation. AAV9 and AAVrh10 have also been tested for GLD therapy in the canine model [266,267]. Combined intravenous and intracerebroventricular delivery of AAVrh10 encoding canine GALC (AAVrh10-cGALC) resulted in delayed symptoms onset, prolonged lifespan, correction of biochemical defects, and also positively affected neuropathology in treated animals [266]. Similarly, AAV9-cGALC has shown promising therapeutic potency in the GLD canine model [267]. Reportedly, intrathecal delivery of AAV9-cGALC resulted in increased activity of the deficient enzyme and, therefore, normal levels of its substrate. It also improved myelination, and decreased inflammation both in the CNS and PNS, which, along with prevention of clinical neurological dysfunction, resulted in a significantly prolonged lifespan of treated dogs, compared to the control group. However, sufficient dosing was found to be critical, as high doses significantly extended the lifespan even for post-symptomatic subjects, and a 5-fold lower dose of the vector resulted in an attenuated form of disease [267].
For ocular diseases, AAV viral vectors have also provided great therapeutic opportunities both in vivo and in human clinical trials [136,170,268,269,270,271]. In a study by Petrs-Silva et al., the efficacy of intraocular transduction of rAAV serotypes 2, 8, and 9 was investigated in mice for targeting ocular neurovascular and retinal diseases [170]. Besides transduction of retinal cells, both serotypes 8 and 9 efficiently transduced the ganglion cell layer, providing evidence for their potentials in therapeutic approaches of ocular diseases. Moreover, AAV2/7 and AAV2/8 hybrid serotypes have achieved high transduction rates of the murine photoreceptors, showing promising potentials for treatment of inherited photoreceptor diseases [136]. AAV2/6 has also shown an efficient transduction of murine cone photoreceptors following subretinal injection [268]. In humans, AAV2/4 has been used in a clinical trial for treatment of childhood blindness caused by Leber Congenital Amaurosis (LCA) disease, mainly by a mutation in the retinal pigment epithelium-specific protein (RPE65) [269,270]. Over a follow-up period between 1 and 3 years, subretinal injection of AAV2/4-RPE65 showed good tolerance both systemically and locally, with no adverse effects, and resulted in an improvement of visual function presented by different parameters in different patients, including improvement of visual acuity and color vision, as well as a reduction of visual fatigue or photophobia [269]. Another example of AAVs’ applications in ocular disease therapy is a phase I/II clinical trial by MacLaren R et al., that has aimed to treat blindness caused by choroideremia, an X-linked recessive disease that results from a mutation affecting retinal escort protein 1 (REP1) [272,273]. In the trial, AAV2-REP1 resulted in improvement of retinal sensitivity in all six patients, following subretinal injection, with two of them having significant increases in visual acuity, supporting further consideration of the tested therapeutic approach [272]. Various other clinical trials have applied AAV-mediated gene therapy for ocular diseases, and many of which have presented promising results, using mainly AAV2 and AAV8 [274,275,276,277,278,279,280,281] (detailed in Table 2).
Similar to ocular diseases, hearing disorders have also been suggested for AAV-mediated gene therapy [282]. In a porcine model, Lalwani A. et al., tested AAV9 intracochlear delivery to assess its efficacy as a vector for gene therapy of hearing disorders, and demonstrated, as a result, transgene expression in the inner ear of animals following administration [282].

5.2. AAV Viral Vectors for Gene Therapy of Respiratory Diseases

For over 20 years, AAV vectors have been vastly investigated for gene therapeutics of respiratory diseases, both in preclinical experiments and human clinical trials [283,284]. However, a key limitation was that many AAV serotypes cannot efficiently transduce airway epithelial cells through the apical surface, suggesting molecular modifications of such serotypes to enhance their transduction efficiency [196]. Cystic fibrosis (CF) represents one of the most studied diseases as a target for AAV-mediated gene therapy [285]. In rabbits, AAV has proven to promote an efficient and stable gene transfer of the cystic fibrosis transmembrane conductance regulator (CFTR) gene into the airway epithelium, indicating, as a result, the vector potential to be used for gene therapy [286]. Subsequently, since 1998, clinical trials have started using recombinant AAV viral vectors to target different sites of the airway epithelium of CF patients, using variant administration routes [285,287,288]. The very first clinical trial to do so used the maxillary sinuses for delivery of recombinant AAV2 containing the CFTR gene (AAV2-CFTR) in ten CF patients, demonstrating a safe, successful transduction of targeted cells and a detected function restoration of the sinuses [288]. Soon after, a following phase II double-blind clinical trial tested unilateral administration of the same vector into the sinus for 23 CF patients, with an in-patient control, achieved by administering a placebo drug into the other sinus [289]. The approach again showed safety and good tolerance, although it did not confirm clinical effectiveness of the treatment. Aitken et al., also tested AAV2-CFTR in a phase I clinical trial for twelve mild CF patients using aerolization by nebulation for delivery, confirming the approaches safety, but failing to yield an effective clinical treatment [290]. Further trials on AAV2-CFTR with single or repeated dosing have yielded similar results of safety and good tolerance, providing little evidence of clinical treatment after intranasal and endobronchial delivery [291,292,293]. In order to optimize such therapeutic approaches and produce a functional CFTR in CF patients, further modifications of recombinant AAV vectors have since been developed [294,295]. Alpha-1 antitrypsin (α1AT or AAT) deficiency is another lung and liver disease that has been extensively studied as a target for AAV-mediated gene therapy both in vivo and in humans [284,296]. Reportedly, intravenous delivery of recombinant AAV vector carrying the human alpha-1 antitrypsin gene (AAV-hAAT) in the murine model resulted in potentially therapeutic serum levels of AAT [297,298]. Further research found that intrapleural delivery of rAAV2-hAAT or rAAV5-hAAT could achieve higher AAT levels both in lungs and serum compared to intramuscular delivery in C57BL/6 mice, with the rAAV5 showing 10-fold higher effectiveness than rAAV2 [299,300]. However, rAAV6/2-hAAT has been shown to transduce murine lung cells even more efficiently than rAAV5 both in vivo (in murine lung cells) and in vitro (in human airway epithelial cell culture) [301]. Similarly, rAAV8-hAAT has also been shown to provide murine lung cells’ transduction superior to that of rAAV5 following intratracheal delivery, as it resulted in 6-fold and 2.5-fold AAT levels in serum and broncho-alveolar fluid, respectively [302], and a high transduction rate of murine hepatocytes [298]. Chiuchiolo M et al., also proved the safety of rhAAV10-hAAT as a viral vector for treatment of AAT in wild-type murine and NHP models [303]. Besides, the delivered vector resulted in persistent expression of the transgene in chest cavity cells of both models, suggesting the efficacy of the tested therapeutic approach [303,304]. Examples of other rAAV vectors’ therapeutic potentials for AAT treatment have been shown in a variety of other preclinical studies, including rAAV1-AAT, rAAV2/9-AAT, and rAAV6 [46,120,175,305].
AAVs have also been tested to develop therapeutic approaches for other lung diseases, such as asthma and surfactant B deficiency [306,307]. A preclinical study by Zavorotinskaya et al., investigated the efficacy of rAAV carrying the interleukin 4 gene (IL-4) as a vector for gene therapy of allergic asthma, upon intratracheal delivery into the murine model [306]. Showing a significant inhibition of airway eosinophilia and mucus production along with a reduction in airway hyper-responsiveness and asthma-associated cytokines levels, obtained data suggested rAAV efficacy for gene therapy of the studied disease. Kang et al., have recently tested an engineered rAAV6/2 vector carrying human or murine surfactant protein B gene (SFTPB) in a murine model [307]. Mutations of this gene cause surfactant protein B deficiency (SPB) in humans, a rare genetic condition with a very poor prognosis [284,308]. In their study, Kang et al., demonstrated efficient transduction of the airway and alveolar epithelium by engineered rAAV6/2-SFTPB following intratracheal administration into a murine model of SPB. As a result, the administered vector was well-tolerated with no adverse effects. Moreover, rapid, long-term restoration of the deficient SPB protein was reported, along with an improvement of lung function, leading, subsequently, to an extended survival [307].

5.3. AAV Viral Vectors for Gene Therapy of Muscle Diseases

Having a large, body-distributed mass, and myofibers with a long half-life time makes muscles an attractive target for gene therapy, also considering the minimal invasiveness of the intramuscular delivery route [133,309]. Muscular dystrophies (MD) constitute a large group of muscle diseases that have been studied for a long time and proved to be a good target for AAV-mediated gene therapy [310]. For example, AAV1 carrying the follistatin gene, an antagonist of muscle growth negative regulator, has been shown to promote sustained improvement in muscle size and strength in NHPs following intramuscular administration [311]. There has also been many reports of both histopathological and functional sustained muscle restoration as a result of using AAV-mediated therapeutic approaches in animal models of Duchenne muscular dystrophy (DMD) [312,313], and Limb-Girdle muscular dystrophies (LGMD) [314,315,316,317,318]. As for clinical research, there have been several registered clinical trials of AAV-mediated gene therapy for muscular dystrophies [319]. The first clinical study was a randomized, double-blind, placebo-controlled phase I clinical trial, conducted in 2012 by Bowles D et al., which used AAV2.5 carrying the mini-dystrophin gene for intramuscular administration in DMD patients [320]. As a result, the administered vector was found to be safe and well-tolerated, and transgene DNA was detected in all patients. Similarly, other clinical trials were started in 2017 and 2018 to test AAV-micro-dystrophin efficacy for treatment of DMD [319] (details in Table 2).

5.4. AAV Viral Vectors for Gene Therapy of Cardiovascular and Blood Diseases

Cardiovascular diseases make another attractive target for gene therapy, being a leading cause of death globally with their high incidence and mortality rates [128,321]. As mentioned before, AAV serotypes 6, 8, and 9 have been shown to efficiently transduce cardiac cells in different animal models, and accordingly, they have been used for therapeutic purposes in different cardiovascular diseases [126,128,167,168,322].
In the search for an effective gene therapy platform for heat failure (HF), White J et al., published a study in 2011, in which they suggested a novel technique for AAV-mediated myocardial gene therapy using molecular cardiac surgery [128]. Using AAV6 in an ovine model, the suggested approach resulted in a global transgene expression, that was cardiac-tropic and substantially more robust and targeted, compared to that of intramuscular or intracoronary injection. Furthermore, a randomized phase I/II AAV1-based clinical trial for heart failure treatment was conducted in 2013, where the researchers used a sarcoplasmic reticulum calcium ATPase gene (SERCA2a), the product of which was suggested to play a key role in HF pathology [323]. As a result, adverse effects, including death, were found to be highest in the placebo group, and lowest in the high-dose group, with evidence of long-term transgene expression. However, in the low-dose and mid-dose groups, adverse effects were also found to be high but delayed.
Hemophilia, an inherited disease that is caused either by deficiency of blood clotting factor VIII (type A) or IX (type B), is another example of diseases that have been targeted by AAV-mediated gene therapy [168,322]. rAAV6 and 8 expressing the canine gene of FVIII have been shown to restore physiologic levels of the deficient factor in a canine model for three years following intravenous administration, without any toxicity or immune reactions [322]. The same vectors could result in a similar effect in the murine model; however, neutralizing antibodies against cFVIII were detected in the mice sera [322]. As for hemophilia B, intravenous administration of rAAV8 and 9 carrying the IX factor gene in a murine model resulted in a significant increase in transgene expression and therefore in IX factor levels, with a decreased proinflammatory risk [168].

5.5. AAV Viral Vectors for Gene Therapy of Liver Diseases

AAV-based gene therapy plays a significant role in liver diseases, as it is, in some cases, an alternative to the only effective therapy, which is liver transplantation [324]. In addition to the previously mentioned diseases affecting the liver, that have been targeted by AAV-base gene therapy, such as CF, AAT deficiency, and hemophilia, there are other liver diseases that have been targeted and continue to be, using AAV8, mainly.
Wilson’s disease (WD) is a rare autosomal recessive disease caused by mutations in the copper transporter gene, ATP7B, resulting in copper accumulation mainly in the liver, along with some other organs [325,326]. A recent in vivo study, conducted by Murillo et al., proved the efficacy of an AAV-based gene therapy for WD in a murine model of the disease [327]. The vector used in the study was AAV8 carrying an optimized short version of the ATP7B gene (AAV8-mini ATP7B) adjusted to the right size, so it could be produced and delivered more efficiently. As a result, intravenous administration of the recombinant vector to 6- and 12-week-old WD mice could restore copper homeostasis, with 20% hepatocyte transduction being sufficient for correction.
Crigler-Najjar syndrome (CNs) type 1 is another autosomal recessive liver disease caused by deficiency of uridine diphosphate glucuronosyltransferase 1A1 (UGT1A1), which leads to severe inherited unconjugated hyperbilirubinemia [328]. There is currently an ongoing clinical trial investigating AAV8-UGT1A1 as a vector for gene therapy in children with CNs [329]; however, as the preclinical data of the trial suggest a negative effect of liver cells’ proliferation on the efficacy of gene therapy in children, Shi et al., suggested to optimize the approach by starting therapy at a specific age and combining it with an immune suppression regimen [330]. In their in vivo study, Shi et al., found that a stable correction was achieved when AAV8-UGT1A1 was administered to CNs rats at the 28th postnatal day, coupled with a rapamycin-based immunosuppression regimen delivered intraperitoneally.

5.6. AAV Viral Vectors for Gene Therapy of Endocrine Disorders

Being a major health issue for humans all around the world, with relatively high prevalence estimates, endocrine disorders have always been in the focus of research for therapeutic approaches and techniques [331,332]. Accordingly, rapid developments of diagnostic techniques, along with the better understanding of endocrine disorders’ pathophysiology and molecular bases, helped in developing different therapies, such as hormonal replacement therapies, for example [332,333]. However, as these approaches have mostly focused on improving the patient’s quality of life, reducing or reversing symptoms, rather than targeting the underlying defect, the resulting effect has not always been sufficient, which highlighted the need for gene therapy approaches, including AAV-based ones [333].
Type 1 diabetes mellitus (T1DM) is one of the endocrine autoimmune disorders that has been extensively studied as a target of gene therapy [334]. T1DM is characterized by self-destruction of insulin-secreting islet β cells, resulting from a wide variety of causative factors [335]. In non-obese diabetic (NOD) mice, administration of recombinant adeno-associated virus serotype 8 carrying DNA of mouse insulin promoter (dsAAV8-mIP) has been shown to prevent hyperglycemia in a dose-dependent manner [336]. High levels of mouse interleukin 10 (mIL-10) achieved following rAAV2-IL-10 intramuscular administration also proved to have a positive effect in NOD mice by decreasing autoimmunity, and thereby hyperglycemia [337]. Similarly, negative regulation of the immune response by programmed death ligand 1 (PDL1) was achieved in NOD mice following intraperitoneal delivery of AAV8-PDL1, which protected β cells [338]. Currently, there is an ongoing clinical trial investigating AAV8 containing a transgene for the fragment antigen-binding region of anti-vascular endothelial growth factor (anti-VEGF fab), as a vector for gene therapy of diabetic retinopathy, delivered in the suprachoroidal space [339].
Autoimmune polyglandular syndrome type-1 (APS1) is another disorder characterized by multiple endocrine abnormalities, resulting from a monogenic defect in the autoimmune regulator (AIRE) gene [340]. Recently, Almaghrabi tested AAV9-AIRE as a potential gene therapy for APS1 [341]. As a result, AAV9-AIRE, following intra-thymic administration in a murine model, showed high transduction efficiency, along with restoration of AIRE expression in the thymus. Subsequently, a significant reduction of serum auto-antibodies was detected in treated mice, with a relatively normal tissue morphology showing no lymphocytic infiltrations.

5.7. AAV Viral Vectors for Gene Therapy of Cancer

Despite the variety of therapeutic approaches developed so far for cancer, including chemotherapy, radiotherapy, surgery, and different medications, cancer is still a huge health issue and a leading cause of death worldwide [342,343]. Several AAV-mediated cancer gene therapy approaches have been reported so far, including suicide gene, RNA-interference, and anti-angiogenesis gene therapies [342].
In AAV-mediated suicide gene therapy, an AAV-based system with herpes simplex virus type-1 thymidine kinase and ganciclovir (AAVtk/GCV) has been used. This system was found to exhibit a significant in vitro and in vivo tumor-suppressor efficacy in human head and neck cancer xenografts in a murine model [344,345,346], and murine models of bile duct cancer [347] and bladder carcinoma [348].
As for AAV-mediated RNA-interference cancer gene therapy, an AAV vector is used to provide a stable expression of short hairpin-structured RNA (shRNA) for gene knockdown applications [349]. This approach could induce a strong androgen receptor (AR) gene silencing in a murine model of prostate cancer following intravenous administration [350]. It has also been tested in a murine model of hepatocellular carcinoma and was found to induce tumor-specific apoptosis, resulting in a significant inhibition of cancer cell proliferation without toxicity [351].
AAV-mediated anti-angiogenesis cancer gene therapy targets factors that can regulate a cancer cell angiogenesis process. VEGF, being an important factor for angiogenesis, has been targeted in a murine model of breast carcinoma using AAV2 carrying transgene of VEGF Trap protein (a pseudo high-affinity receptor) [352]. The approach resulted in significant suppression of tumor growth and prevention of spontaneous pulmonary metastases. Another factor that has been targeted by this approach is pigment epithelium-derived factor (PEDF), a highly potent angiogenesis inhibitor. In a murine colorectal peritoneal carcinomatosis model, AAV2 carrying the human gene of PEDF was found to induce significant tumor suppression, inhibit metastases, and prolong survival time [353]. Similarly, kallistatin is another anti-angiogenesis factor that has been targeted using an AAV-mediated murine model of hepatocellular carcinoma, and it provided a similar efficacy as AAV-kallistatin-induced potent cancer cell apoptosis [354].

6. Challenges and Limitations of AAVs for Gene Therapy AAV-Mediated Applications

Despite the many advantages AAVs provide as viral vectors for gene therapy of various diseases, and the many improvements and molecular modifications they have undergone since their first discovery, AAVs still have some limitations that can result in impeding impacts on their applications in some cases. One example of such major limitations is the limited cloning capacity of AAVs, that is ~4.5 Kb [355,356]. Moreover, AAVs need to converse into double-stranded DNA, and may cause a delay in transgene expression following delivery, compared to double-stranded vectors [357]. Although it can be dose-dependent in some cases, preexisting immunity against some serotypes can also be a limiting factor [358,359].
Preexisting immunity, or an immunity response occurring after the first introduction of a viral vector, can be a major obstacle in the development of gene therapy approaches based on AAV or other viral vectors, and their application in clinical practice. Neutralizing antibodies are often found in the human population with a high prevalence that varies depending on the virus serotype and the geographic location of the population. Moreover, in a number of diseases, multiple administration of an AAV-based drug may be required, which can result in developing an immune response against the viral vector. Neutralizing antibodies can have a significant impact on the efficiency of gene transfer in gene therapy. Consequently, many patients with genetic diseases will not be able to benefit from such gene therapy medications.
Furthermore, although AAVs have not been linked to human pathologies or diseases, wild-type AAVs are widely distributed in the human population, and despite having a low immunogenicity profile, an immune response was observed in humans as a result of gene therapy using AAVs [358]. To date, the impact of such anti-AAV immune response on gene therapy is still not entirely understood, however, its mechanisms are being actively explored. For example, systemic delivery of AAV vectors, while having neutralizing antibodies, has been shown to result in accumulation of vector genomes in lymphoid organs. Whereas the efficiency of liver transduction while having binding antibodies (indicative of a previous infection, but not neutralizing the virus), on the contrary, increases [360]. The role of neutralizing antibodies in activating the complement system has also been studied [361]. Individuals seropositive for AAV have high titers of IgG1, IgG2, and IgG3. It is known that IgG1, IgG2, and IgM highly correlate with the level of neutralizing antibodies, and IgG1 levels increase after gene therapy with AAV [362]. Additionally, IgG3 correlates with the level of T-cell reactivity against AAV vectors [363]. From a clinical point of view, neutralizing antibody titers are of a great significance, as having neutralizing antibodies against a specific AAV serotype highly affects gene therapy effectiveness upon using this serotype. Although methods for detecting neutralizing antibodies are poorly standardized, existing research data indicate that the prevalence of neutralizing antibodies against AAV2 is about 40% [364]. It has been shown that neutralizing antibodies can trigger elimination of the vector by the immune system. At the same time, viral vectors can accumulate in the spleen, not reaching the liver [365]. Therefore, researching mechanisms to establish and develop effective methods for preexisting immunity assessment can help overcome many problems arising in gene therapy applications. Tackling such an urgent problem can expand the possibilities of therapy for many incurable human hereditary diseases [366].
Another challenge is the need to deeply understand the exact mechanisms and pathways of AAV cellular uptake, trafficking, and transduction, which are still unknown for most AAV serotypes [367]. Such understanding can help in addressing and explaining the changes of transduction potency and efficacy that can be found between different species in different in vivo studies, or between treatment efficacy in vivo and in clinical trials [368]. In other words, differences between humans and various animal models in terms of body size, genomic and anatomic specificity, immunogenicity, and disease process should be taken into account, as they can affect the effective translation of preclinical investigations [369].
Table 1. Characteristics of different AAV serotypes.
Table 1. Characteristics of different AAV serotypes.
AAV SerotypePrimary ReceptorOther/CoreceptorsPost Translational ModificationsRecommended Purification MethodTropism
AAV1Sialic acidAAV receptor (AAVR)-iodixanol gradient centrifugation, anion-exchange chromatography, and mucin column affinity chromatographySkeletal muscles [35,43,44,45,46], heart [48,49], glial and ependymal cells in the murine brain [47], endothelial and vascular smooth muscles [50], retina [51]
AAV2HSPGFGFR1, αVβ5 and α5β1 integrins, HGFR, LR, and CD9ubiquitination, phosphorylation, SUMOylation, and multiple-site-glycosylationheparin column affinity chromatographyrenal tissue [62,63,64], hepatocytes [65,66], retina [67,68,69], non-mitotic cells of central nervous system (CNS) [70,71], and skeletal muscles [45,71]
AAV3HSPGFGFR1, LR, and HGFRacetylation, phosphorylation, and glycosylationiodixanol gradient centrifugationhuman liver cancer cells as well as human and NHP hepatocytes [76,77], murine cochlear inner hair cells [82].
AAV4Sialic acid-ubiquitinationion-exchange chromatography, mucin column affinity chromatographyependymal cells of mammalian CNS [90], RPE cells of the retina (canine, rodent, and NHP origins) [91], murine kidney, lung, and heart cells [92,99].
AAV5Sialic acidPDGFR (α and β)ubiquitination, phosphorylation, SUMOylation, and glycosylationion-exchange chromatography, mucin column affinity chromatographymurine: retinal cells [38,51], mainly photoreceptors [106], airway epithelia [98,110], liver cells [92,112], vascular endothelial cells and smooth muscles [50], and neurons (murine and NHP) [107,370]
AAV6Sialic acid and HSPGEGFRacetylationheparin or mucin column affinity chromatographyairway epithelia of murine and canine models [119,120], murine liver cells [92,121], skeletal muscles of murine and canine models [92,122,123,124], cardiomyocytes in murine [92,125], porcine [126], canine [127], and in sheep [128] models.
AAV7--glycosylation, phosphorylation, SUMOylation, and acetylation-murine skeletal muscle cells [133], murine and human hepatocytes [92], murine and NHPs CNS [134,135], murine photoreceptor cells [136], murine vascular endothelial cells (limited tropism) [130], murine epicardium cells [49].
AAV8LR-phosphorylation, glycosylation, and acetylationDual-ion-exchange chromatography,
iodixanol gradient centrifugation
murine, canine, and 115 hepatocytes [92,112,141,142,143,144,145,146,147], murine skeletal and cardiac muscles [148], murine pancreatic cells [149,150], murine renal cells [151], and different cells in the murine retina [152,153,154]
AAV9terminal N-linked galactoseputative integrin, LRglycosylation, ubiquitination, phosphorylation, SUMOylation, and acetylationSucrose gradient centrifugation, and ion-exchange chromatographymurine, NHP, and feline neuronal and non-neuronal cells, including astrocytes [27,159,160,161,162], murine and NHP retinal photoreceptors cells [163], murine, NHP, and porcine cardiac muscle tissue [164,165,166,167,168,169], murine hepatocytes, skeletal muscles, and pancreatic cells [81,169], photoreceptor cells [170], renal tubular epithelium cells [171,172,173], Leydig cells in the testicular interstitial tissue [174], and alveolar and nasal epithelia [175,176].
AAV10--glycosylation, ubiquitination, phosphorylation, SUMOylation, and acetylationiodixanol gradient centrifugationNHP intestinal cells, hepatocytes, lymph nodes, and less frequently, renal cells and adrenal glands [180], murine small intestine and colon cells [181], retinal cells (including RPE, cells in the ganglion cell layer, several cell types in the inner nuclear layer, photoreceptors, and a highly efficient transduction of horizontal cells) [152], murine liver cells, lung cells [182], renal, and pancreatic cells [183].
AAV11---iodixanol gradient centrifugationNHP intestinal cells, hepatocytes, lymph nodes, and less frequently, renal cells and adrenal glands [180], murine projection neurons and astrocytes [184], and mild tropism to NHP CNS (cerebrum and spinal cord, mainly) [180].
AAV12mannose and mannosamine have been suggested as components of a potential receptor complex-AVB Sepharose affinity chromatographymurine salivary glands and muscles [185], murine nasal epithelia (mainly after intranasal administration) [187].
AAV13-HSPG-iodixanol gradient centrifugation,
heparin column affinity chromatography
-
Table 2. Preclinical and clinical studies using AAVs as viral vectors for gene therapy of different diseases.
Table 2. Preclinical and clinical studies using AAVs as viral vectors for gene therapy of different diseases.
Gene Therapy TargetDisease **AAV Viral VectorStudy TypeOutcomeRef.
CNSPDAAV2-GADphase I clinical trial for advanced PD patientssafe and well-tolerated approach, providing significant improvement in motor function scores up to 12 months after unilateral subthalamic delivery[229]
double-blind, controlled, randomized clinical trial for advanced PD patientssafe and well-tolerated approach, along with improved motor function scores following bilateral subthalamic delivery[230]
AAV2-neurturinopen-label clinical trial for PD patientssuggested feasibility, safety, and good tolerance of the approach[231]
rAAV2-AADCclinical trial for moderately advanced PD patientsgood tolerance, improvement on motor rating scales, however, accompanied with an increased risk of intracranial hemorrhages and headache [232]
good tolerance and a stable expression of the transgene that lasted for the following 4 years, although higher vector doses were suggested for further studies[233]
AADC deficiencyrAAV2-AADCopen-label, phase I/II trial in childrengood tolerance in general, with evidence for potential improvement of motor function[234]
SMAAA9-SMNopen-label, phase I clinical trialsignificant improvement of the motor function in all 15 patients following single-dose intravenous administration, reflected by their ability to perform different activities, such as unassisted sitting and walking, oral feeding, and speaking, with no reported motor function regression at two-year follow-up. A long-term safety assessment, however, was recommended.[236]
LSDMPS VII(rAAV-GUSβ)in vivo study, murine modelstable expression of the deficient enzyme upon single administration, that was adequate for phenotype correction in the liver[238,239]
(rAAV-GUSβ)in vivo study, neonatal murine modeltherapeutic levels of the deficient enzyme in multiple organs, including ones of the CNS, with the gene expression not being affected by rapid growth and differentiation of tissues[240]
rAAV9-GUSβ and
rAAVrh10-GUSβ
in vivo study, canine modelsignificantly high expression levels of GUSβ in the CNS tissues, with the enzyme in brain tissue homogenates showing over 100% normal activity [221]
MPS IIIAAAV9-Sgsh (canine sulfamidase gene)in vivo study, canine modelsustained and widely distributed transgene expression with no toxicity for a duration of 7 years after therapy[241]
MLDAAV1-ARSAin vivo study, murine modelsignificant elevation of ARSA levels and activity, resulting in reduction of accumulated sulfatides[244]
AAV5-ARSAin vivo study, murine modelrapid, abundant, and sustained restoration of ARSA levels in the brain and brainstem up to 15 months after administration, reduction of accumulated sulfatides, and preservation of neurologic function.[246]
AAV5-ARSAin vivo study, NHP modelgood tolerance, distribution of the transgene in the brain with elevated activity of the deficient enzyme[370]
AAV9-ARSAin vivo study, neonatal murine modelglobal expression of the transgene in the brain and spinal cord, along with muscles and heart, inhibition of sulfatide accumulation, and improvement of neurologic/motor function[245]
AAVrh10-ARSA
AAVrh10-ARSA
AAVrh10-ARSA
in vivo study, murine modelcorrection of sulfatide accumulation following single administration, with a transduction efficacy higher than that of AAV5 as it transduced both neurons and oligodendrocytes[243]
in vivo study, NHP modelgood tolerance, neuroinflammation 3 months following the fifth dose but none after the first, detection of transgene expression after the first dose along with increased ARSA activity, and detection of the enzyme in other organs but not in gonads[371]
clinical trial for children with asymptomatic or early-stage MLDsignificant elevation in ARSA levels in the cerebrospinal fluid (CSF) following intracerebral vector delivery, but no clinical improvement has been noticed compared to the control group.[253]
GM2-gangliosidoserAAV2-HEXA + rAAV2-HEXBin vivo study, murine modeldelay of disease clinical onset, maintained motor function, good tolerance, stable and abundant levels of the deficient enzyme, resulting in a significant reduction of gangliosides’ storage in the CNS[250]
AAVrh8-HEXin vivo study, feline modelprevention or reduction of tremors that is characteristic of improvement in the neurologic function, and a signification elevation of the deficient enzyme levels (HEX), resulting in a significant reduction of gangliosides’ storage in different tissues of the CNS.[252]
AAVrh8-HEXA +
AAVrh8-HEXB
in vivo study, ovine modeldelay of disease clinical onset and progression, improved neurologic function and clinical biomarkers. However, lifespan was not significantly higher.[251]
AAVrh8-HEXA +
AAVrh8-HEXB
first clinical trial for children with TSDgood tolerance and broad distribution of the transgene in the CNS[248]
CDrAAV2-ASPAin vivo study, murine modelincreased ASPA activity and, thereby, decreased NAA accumulation and white matter degeneration. However, areas remote from injection site, such as the cerebellum, were not affected[256]
GLDrAAV1-GALCin vivo study, murine modelsustained expression of the deficient enzyme, improved myelination status, and prolonged lifespan. However, both treated and untreated mice died with similar symptoms, suggesting that the used approach should be initiated prior to symptoms’ onset.[262]
rAAV2/5-GALCin vivo study, murine modelwide dispersion of GALC transgene across the CNS reaching areas remote from the injection site, reduced loss of oligodendrocytes and Purkinje cells, along with a significant improvement of neuromotor function and a prolonged lifespan of treated mice.[263]
AAV9-GALC
AAVrh10-GALC
AAVOlig001-GALC
in vivo study, murine modelAll three serotypes provided wide distribution of the transgene across the CNS and liver, resulting in a significant improvement of myelination, and a prolonged lifespan, with AAV9 being the most effective when combined with bone marrow transplantation.[265]
AAVrh10-cGALCin vivo study, canine modeldelayed symptoms onset, prolonged lifespan, correction of biochemical defects, and a positive effect on neuropathology in treated animals [266].
AAV9-cGALCin vivo study, canine modelincreased activity of the deficient enzyme and, therefore, normal levels of its substrate, improved myelination, and decreased inflammation both in the CNS and PNS, which, along with prevention of clinical neurological dysfunction, resulted in a significantly prolonged lifespan of treated dogs, compared to the control group. However, sufficient dosing was found to be critical, as high doses significantly extended the lifespan even for post-symptomatic subjects, and a 5-fold lower dose of the vector resulted in an attenuated form of disease[267].
Retinal degeneration and ocular neurovascular diseasesAAV2, 8, and 9in vivo study, C57BL/6 micetransduction of retinal cells, as well as efficient transduction of ganglion cell layer by AAV8 and AAV9[170]
Inherited photoreceptor diseaseshybrid serotypes AAV2/7 and AAV2/8in vivo study, C57BL/6 micehigh transduction rates of the murine photoreceptors[136]
Retinal blindness caused by LCAAAV2/6in vivo study, wild-type 129Sv/Ev
(Taconic) mice
efficient transduction of murine cone photoreceptors following subretinal injection[268]
AAV2/4- RPE65clinical trial for patients with LCAsystemic and local good tolerance with no adverse effects, along with improvement of visual function presented by different parameters in different patients, including improvement of visual acuity and color vision, as well as reduction of visual fatigue or photophobia, over a follow-up period between 1 and 3 years
[269]
Retinal blindness caused by choroideremiaAAV2- REP1phase I/II clinical trialimprovement of retinal sensitivity in all six patients, following subretinal vector delivery, with two of them having significant increases in visual acuity, supporting further consideration of the tested therapeutic approach.[272]
RPE65-mediated inherited retinal dystrophyAAV2-hRPE65v2open-label, randomized, controlled phase III trialgood tolerance with no adverse effects, restoration of
RPE65 enzymatic activity reflected by significant and sustained improvement in light perception and navigational abilities
[274]
Leber Hereditary Optic Neuropathy (LHON)rAAV2-ND4
(gene encoding nicotinamide adenine dinucleotide
dehydrogenase subunit IV)
open-label, phase I/II randomized clinical trialgood tolerance, although a mild, intraocular inflammation was detected after vector administration, but it was responsive to treatment and suggested to be overcome in further studies by systemic vector delivery instead of local.[275]
open-label, phase I clinical trialminor adverse events, improvement in visual activity in some but not all patients, no detection of vector DNA in patients’ blood samples[276]
RPGR-related X-linked retinitis pigmentosa AAV8-RPGRphase I/II clinical trialno adverse effects other than steroid-responsive subretinal inflammation following administration of higher doses, sustained improvements in visual function in 6/18 patients.[277]
X-Linked RetinoschisisAAV8-RS1
(Retinoschisin gene)
phase I/IIa single-center, open-label, clinical trialgood tolerance in general, although steroid-responsive, dose-related inflammation was observed, and a dose-related increase of systemic antibodies against AAV8, but none against RS1[278,279]
CNGA3-linked achromatopsiaAAV8-CNGA3nonrandomized controlled clinical trialno substantial safety issues, successful targeting of cone photoreceptors reflected by reported improvement of color vision ability, along with improvements in visual acuity and contrast sensitivity (although a cause–effect relationship was not established) [280,281]
Hearing disordersAAV9in vivo study, porcine modelpersistent expression of the transgene within the mammalian inner ear following intracochlear delivery[282]
Respiratory organs (airway epithelia)CFAAV2-CFTRin vivo study-rabbitsefficient and stable gene transfer of CFTR into airway epithelium, indicating, as a result, the vector potential to be used for gene therapy [286]
phase I clinical trial for CF patientssafety, successful transduction of targeted cells, and a detected function restoration of sinuses.[288]
phase II, double-blind, randomized, placebo-controlled clinical trial for CF patientssafety and good tolerance, but no effective clinical treatment of disease was achieved[289]
phase I clinical trial for mild CF patientssafety, but no effective clinical treatment of disease was achieved[290]
α1AT deficiencyrAAV-hAATin vivo study, intravenous delivery into murine modelSerum levels of AAT that are potentially therapeutic[297,298]
rAAV2-hAAT and rAAV5-hAATin vivo study, intrapleural vs. intramuscular delivery into murine modelAAT lung and serum levels higher than the ones achieved by intravenous delivery, with rAAV5 showing 10-fold higher effectiveness than rAAV2[299,300]
rhAAV10-hAATin vivo study, intrapleural delivery into murine and NHP modelssafety, and persistent expression of the transgene in the chest cavity cells of both models[303,304]
rAAV6/2-hAATin vivo study, intratracheal delivery into C57/Bl6 mice,
and in vitro study, cultures of human airway epithelial cells
lung cells’ transduction, even more efficient than rAAV5 both in vivo and in vitro[301]
rAAV8-hAATin vivo study, intratracheal delivery into C57/Bl6 micelung cells’ transduction superior to that of rAAV5, as it resulted in 6-fold and 2.5-fold higher AAT levels in serum and broncho-alveolar fluid, respectively.[302]
rAAV2/8-hAATin vivo study, intravenous delivery into C57/Bl6 micehigh transduction rate of hepatocytes[298,372]
rAAV1-AATin vivo study, intramuscular delivery into C57/Bl6 mice and rabbitsdose-dependent inflammatory infiltrates at injection sites not affecting expression of transgene, along with dose-dependent detection of vector DNA in most animals, both at injection sites and in distal organs[46,305]
rAAV2/9-AATin vivo study, intratracheal delivery into C57/Bl6 miceefficient, relatively stable transduction of alveolar and nasal epithelia, that was not affected by high levels of neutralizing antibodies upon following re-administration [175]
rAAV6-hAATin vivo study, nasal delivery into C57/Bl6 mice, and intratracheal delivery into canine modeltherapeutic hAAT concentrations in both studied animal models, with levels being higher in lungs than serum, accompanied with an immune response against the vector capsid in some animals despite being immunosuppressed[120]
Allergic asthmarAAV-IL-4in vivo study, intratracheal delivery into Balb/cByJ micesignificant inhibition of airway eosinophilia and mucus production along with a reduction in airway hyper-responsiveness and asthma-associated cytokine levels[306]
SPBrAAV6/2-SFTPBin vivo study, intratracheal delivery into murine modelefficient transduction of airway and alveolar epithelium, good tolerance of administered vector with no adverse effects, rapid and long-term restoration of the deficient SPB protein, along with an improvement of lung function, leading, subsequently, to an extended survival[307]
MusclesDegenerative muscle disordersAAV1-FS344
(follistatin gene)
in vivo study, NHP modelsafety and good tolerance of administered vector, promotion of sustained improvement in muscle size and strength[311]
DMDrAAV6-micro-dystrophinin vivo study, dystrophin/utrophin double-knockout murine modelno serious adverse events, improvement of muscle function along with prolongation of lifespan, explained by sustained restoration of deficient protein (dystrophin) in the respiratory, cardiac, and limb muscles[312]
rAAV1-mini-dystrophinin vivo study, dystrophin/utrophin double knockout murine modelprolongation of lifespan, highly efficient expression of transgene, improvement in muscle histopathology and function, reflected by improved growth and motility, along with
prevention of spine and extremities’ deformation
[313]
AAV2.5-mini-dystrophinRandomized, double-blind, placebo-controlled phase I clinical trialsafety and good tolerance of the vector, and detection of transgene DNA in all patients following intramuscular administration[320]
AAV9-micro-dystrophinphase I/II open-label, randomized, controlled clinical trialsafety and successful expression of transgene[319]
AAV-rh74-micro-dystrophinphase I/II, open-label, non-randomized clinical trialsafety and successful expression of transgene, along with improvement of motor function
LGMDrAAV8-hδ-SG (human δ-sarcoglycan gene)in vivo study, murine modelprolongation of lifespan, efficient and sustained transduction of cardiac and skeletal muscles along with improvement in their histopathology and function[314]
rAAV1-hα-SGin vivo study, murine modelefficient and sustained expression of the transgene, histological and functional improvement of musculature reflected by correction of contractile force deficits and stretch sensibility, along with increase of animal activity[316]
rAAV-hγ-SGin vivo study, murine modelsignificantly efficient transduction of muscle fibers, and general histopathological improvement, that were achieved only upon early intervention[317]
Cardiovascular systemHFAAV6-EGFPin vivo study, ovine modelglobal expression of the transgene, that was cardiac-tropic and substantially more robust and targeted, compared to that of intramuscular or intracoronary injection.[128]
AAV1-SERCA2arandomized phase I/II clinical trialhighest adverse effects, including death, in the placebo group, and lowest in the high-dose group, with evidence of long-term transgene expression. However, in the low-dose and mid-dose groups, adverse effects were found to be high but delayed.[323]
Hemophilia ArAAV6-cFVIII and rAAV8-cFVIIIin vivo study, murine and canine models restoration of physiologic levels of the deficient factor in the canine model for three years following intravenous administration, without any toxicity or immune reactions.
A similar effect in the murine model was found, however, with detection of neutralizing antibodies against cFVIII in the mice sera
[322]
Hemophilia BrAAV8-FIX and rAAV9-FIXin vivo study, murine modelsignificant increase in transgene expression and therefore in IX factor levels, with decreased proinflammatory risk following intravenous administration[168]
LiverWDAAV8-mini ATP7Bin vivo study, murine modelrestoration of copper homeostasis, with 20% hepatocyte transduction being sufficient for correction[327]
CNsAAV8- UGT1A1ongoing clinical trial-[329]
in vivo study, murine modelstable correction was achieved when the vector was administered at the 28th postnatal day, coupled with a rapamycin-based immunosuppression regimen delivered intraperitoneally[330]
Endocrine systemT1DMdsAAV8-mIPin vivo study, murine modelprevention of hyperglycemia in a dose-dependent manner[336]
rAAV2-IL-10in vivo study, murine modela positive effect, decreasing autoimmunity, and thereby hyperglycemia [337]
AAV8-PDL1in vivo study, murine modelpancreatic β cells’ protection[338]
AAV8-anti-VEGF fabongoing clinical trial for diabetic retinopathy patients-[339]
APS1AAV9-AIREin vivo study, murine modelhigh transduction efficiency, along with restoration of AIRE expression in the thymus, following intra-thymic administration, and a subsequent significant reduction of serum auto-antibodies, with relatively normal tissue morphology showing no lymphocytic infiltrations.[341]
Cancerlaryngeal cancer cell line (HEp-2)AAVtk/GCV systemin vitro and in vivo study, murine modelsignificant tumor-suppressor efficacy in human head and neck cancer xenografts, and markedly prolonged survival[344,345,346]
bile duct cancerAAVtk/GCV systemin vivo study, murine modelincreased anti-tumor effect upon simultaneous administration with 5-fluorouracil, and better survival[347]
bladder carcinomaAAVtk/GCV systemin vivo study, murine modelcontrol of tumor cell growth, a strong anti-tumor efficacy [348]
prostate cancerAAV2-ARHP8in vivo study, murine modelstrong androgen receptor (AR) gene silencing following intravenous administration[350]
hepatocellular carcinomaAAV8-miR-26ain vivo study, murine modelInduction of tumor-specific apoptosis, resulting in a significant inhibition of cancer cell proliferation without toxicity [351]
breast carcinomaAAV2-VEGF Trapin vivo study, murine modelsignificant suppression of tumor growth and prevention of spontaneous pulmonary metastases[352]
colorectal peritoneal carcinomatosisAAV2-hPEDFin vivo study, murine modelsignificant tumor suppression, inhibition of metastases, and prolonged survival time[353]
hepatocellular carcinomaAAV-kallistatinin vivo study, murine modelinduced potent cancer cell apoptosis, and prolonged survival time[354]
** Disease abbreviations: PD: Parkinson’s disease, AADC deficiency: amino acid decarboxylase deficiency, SMA: spinal muscular atrophy, MPS: mucopolysaccharidosis, MLD: metachromatic leukodystrophy, CD: Canavan’s disease, GLD: globoid cell leukodystrophy, LCA: Leber congenital amaurosis, CF: cystic fibrosis, α1AT deficiency: Alpha-1 antitrypsin deficiency, SPB deficiency: surfactant protein B deficiency, DMD: Duchenne muscular dystrophy, LGMD: Limb-Girdle muscular dystrophies, HF: heart failure, WD: Wilson’s disease, CNs: Crigler-Najjar syndrome, T1DM: Type 1 diabetes mellitus, APS1: autoimmune polyglandular syndrome type-1.

7. Conclusions

AAVs were discovered over five decades ago and have since represented a potent tool for gene therapy, that still needs to be better understood and developed for broader and larger therapeutic applications. Further optimizations should cover vector design, tropism modifications, and delivery routes. In this article, we have covered different serotypes of adeno-associated viruses and their applications in gene therapy for different diseases in preclinical and clinical studies, with a brief overview of AAVs’ limitations and challenges.

Author Contributions

Conceptualization, S.S.I., A.A.S. and V.V.S.; writing—original draft preparation, S.S.I. and A.A.S.; writing—review and editing, V.V.S.; supervision, A.A.R. All authors have read and agreed to the published version of the manuscript.

Funding

This article has been supported by the Kazan Federal University Strategic Academic Leadership Program (PRIORITY-2030).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We thank the Center for Precision Genome Editing and Genetic Technologies for Biomedicine, IGB RAS, for scientific advice on development of gene therapy drugs.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ozelo, M.C.; Mahlangu, J.; Pasi, K.J.; Giermasz, A.; Leavitt, A.D.; Laffan, M.; Symington, E.; Quon, D.V.; Wang, J.D.; Peerlinck, K.; et al. Valoctocogene Roxaparvovec Gene Therapy for Hemophilia A. N. Engl. J. Med. 2022, 386, 1013–1025. [Google Scholar] [CrossRef] [PubMed]
  2. Zolgensma—One-time gene therapy for spinal muscular atrophy. Med. Lett. Drugs Ther. 2019, 61, 113–114.
  3. Prado, D.A.; Acosta-Acero, M.; Maldonado, R.S. Gene therapy beyond luxturna: A new horizon of the treatment for inherited retinal disease. Curr. Opin. Ophthalmol. 2020, 31, 147–154. [Google Scholar] [CrossRef] [PubMed]
  4. Keeler, A.M.; Flotte, T.R. Recombinant Adeno-Associated Virus Gene Therapy in Light of Luxturna (and Zolgensma and Glybera): Where Are We, and How Did We Get Here? Annu. Rev. Virol. 2019, 6, 601–621. [Google Scholar] [CrossRef]
  5. Atchison, R.W.; Casto, B.C.; Hammon, W.M. Adenovirus-associated defective virus particles. Science 1965, 149, 754–756. [Google Scholar] [CrossRef]
  6. Mezzina, M.; Merten, O.W. Adeno-associated viruses. Methods Mol. Biol. 2011, 737, 211–234. [Google Scholar] [CrossRef]
  7. Hoggan, M.D.; Blacklow, N.R.; Rowe, W.P. Studies of small DNA viruses found in various adenovirus preparations: Physical, biological, and immunological characteristics. Proc. Natl. Acad. Sci. USA 1966, 55, 1467–1474. [Google Scholar] [CrossRef] [Green Version]
  8. Berns, K.I.; Giraud, C. Biology of adeno-associated virus. Curr. Top. Microbiol. Immunol. 1996, 218, 1–23. [Google Scholar] [CrossRef]
  9. Yalkinoglu, A.O.; Heilbronn, R.; Bürkle, A.; Schlehofer, J.R.; zur Hausen, H. DNA amplification of adeno-associated virus as a response to cellular genotoxic stress. Cancer Res. 1988, 48, 3123–3129. [Google Scholar]
  10. Gonçalves, M.A.F.V. Adeno-associated virus: From defective virus to effective vector. Virol. J. 2005, 2, 43. [Google Scholar] [CrossRef] [Green Version]
  11. Erles, K.; Sebökovà, P.; Schlehofer, J.R. Update on the prevalence of serum antibodies (IgG and IgM) to adeno-associated virus (AAV). J. Med. Virol. 1999, 59, 406–411. [Google Scholar] [CrossRef]
  12. Büning, H.; Perabo, L.; Coutelle, O.; Quadt-Humme, S.; Hallek, M. Recent developments in adeno-associated virus vector technology. J. Gene Med. 2008, 10, 717–733. [Google Scholar] [CrossRef]
  13. Tijssen, P. Handbook of Parvoviruses Vol. I and Vol. II Continued; CRC Press, Inc.: Boca Raton, FL, USA, 1990. [Google Scholar]
  14. Grimes, T.; King, D.; Fletcher, O. Serologic and pathogenicity studies of avian adenovirus isolated from chickens with inclusion body hepatitis. Avian Dis. 1978, 22, 177–180. [Google Scholar] [CrossRef]
  15. Alvarado, I.; Villegas, P.; El-Attrache, J.; Jensen, E.; Rosales, G.; Perozo, F.; Purvis, L.J. Genetic characterization, pathogenicity, and protection studies with an avian adenovirus isolate associated with inclusion body hepatitis. Avian Dis. 2007, 51, 27–32. [Google Scholar] [CrossRef]
  16. Bossis, I.; Chiorini, J.A. Cloning of an avian adeno-associated virus (AAAV) and generation of recombinant AAAV particles. J. Virol. 2003, 77, 6799–6810. [Google Scholar] [CrossRef] [Green Version]
  17. Arbetman, A.E.; Lochrie, M.; Zhou, S.; Wellman, J.; Scallan, C.; Doroudchi, M.M.; Randlev, B.; Patarroyo-White, S.; Liu, T.; Smith, P. Novel caprine adeno-associated virus (AAV) capsid (AAV-Go. 1) is closely related to the primate AAV-5 and has unique tropism and neutralization properties. J. Virol. 2005, 79, 15238–15245. [Google Scholar] [CrossRef] [Green Version]
  18. Olson, E.J.; Haskell, S.R.; Frank, R.K.; Lehmkuhl, H.D.; Hobbs, L.A.; Warg, J.V.; Landgraf, J.G.; Wünschmann, A. Isolation of an adenovirus and an adeno-associated virus from goat kids with enteritis. J. Vet.-Diagn. Investig. 2004, 16, 461–464. [Google Scholar] [CrossRef] [Green Version]
  19. Schmidt, M.; Katano, H.; Bossis, I.; Chiorini, J.A. Cloning and characterization of a bovine adeno-associated virus. J. Virol. 2004, 78, 6509–6516. [Google Scholar] [CrossRef] [Green Version]
  20. Coria, M.; Lehmkuhl, H. Isolation and identification of a bovine adenovirus type 3 with an adenovirus-associated virus. J. Vet.-Diagn. Investig. 1978, 39, 1904–1906. [Google Scholar]
  21. Dutta, S.K. Isolation and characterization of an adenovirus and isolation of its adenovirus-associated virus in cell culture from foals with respiratory tract disease. Am. J. Vet. Res. 1975, 36, 247–250. [Google Scholar]
  22. Qiu, J.; Pintel, D.J. Alternative polyadenylation of adeno-associated virus type 5 RNA within an internal intron is governed by the distance between the promoter and the intron and is inhibited by U1 small nuclear RNP binding to the intervening donor. J. Biol. Chem. 2004, 279, 14889–14898. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Merten, O.W.; Gény-Fiamma, C.; Douar, A.M. Current issues in adeno-associated viral vector production. Gene Ther. 2005, 12 (Suppl. S1), S51–S61. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Potter, M.; Lins, B.; Mietzsch, M.; Heilbronn, R.; Van Vliet, K.; Chipman, P.; Agbandje-McKenna, M.; Cleaver, B.D.; Clément, N.; Byrne, B.J.; et al. A simplified purification protocol for recombinant adeno-associated virus vectors. Mol. Ther. Methods Clin. Dev. 2014, 1, 14034. [Google Scholar] [CrossRef] [PubMed]
  25. Carter, P.J.; Samulski, R.J. Adeno-associated viral vectors as gene delivery vehicles. Int. J. Mol. Med. 2000, 6, 17–27. [Google Scholar] [CrossRef] [PubMed]
  26. Cataldi, M.P.; McCarty, D.M. Hairpin-end conformation of adeno-associated virus genome determines interactions with DNA-repair pathways. Gene Ther. 2013, 20, 686–693. [Google Scholar] [CrossRef] [Green Version]
  27. Michelfelder, S.; Trepel, M. Adeno-associated viral vectors and their redirection to cell-type specific receptors. Adv. Genet. 2009, 67, 29–60. [Google Scholar] [CrossRef]
  28. Dubielzig, R.; King, J.A.; Weger, S.; Kern, A.; Kleinschmidt, J.A. Adeno-associated virus type 2 protein interactions: Formation of pre-encapsidation complexes. J. Virol. 1999, 73, 8989–8998. [Google Scholar] [CrossRef] [Green Version]
  29. Kyöstiö, S.R.; Owens, R.A.; Weitzman, M.D.; Antoni, B.A.; Chejanovsky, N.; Carter, B.J. Analysis of adeno-associated virus (AAV) wild-type and mutant Rep proteins for their abilities to negatively regulate AAV p5 and p19 mRNA levels. J. Virol. 1994, 68, 2947–2957. [Google Scholar] [CrossRef] [Green Version]
  30. Galibert, L.; Hyvönen, A.; Eriksson, R.A.E.; Mattola, S.; Aho, V.; Salminen, S.; Albers, J.D.; Peltola, S.K.; Weman, S.; Nieminen, T.; et al. Functional roles of the membrane-associated AAV protein MAAP. Sci. Rep. 2021, 11, 21698. [Google Scholar] [CrossRef]
  31. Kronenberg, S.; Kleinschmidt, J.A.; Böttcher, B. Electron cryo-microscopy and image reconstruction of adeno-associated virus type 2 empty capsids. EMBO Rep. 2001, 2, 997–1002. [Google Scholar] [CrossRef] [Green Version]
  32. Rabinowitz, J.E.; Samulski, R.J. Building a better vector: The manipulation of AAV virions. Virology 2000, 278, 301–308. [Google Scholar] [CrossRef] [Green Version]
  33. Wu, Z.; Asokan, A.; Samulski, R.J. Adeno-associated virus serotypes: Vector toolkit for human gene therapy. Mol. Ther. J. Am. Soc. Gene Ther. 2006, 14, 316–327. [Google Scholar] [CrossRef]
  34. Mori, S.; Wang, L.; Takeuchi, T.; Kanda, T. Two novel adeno-associated viruses from cynomolgus monkey: Pseudotyping characterization of capsid protein. Virology 2004, 330, 375–383. [Google Scholar] [CrossRef] [Green Version]
  35. Gao, G.P.; Alvira, M.R.; Wang, L.; Calcedo, R.; Johnston, J.; Wilson, J.M. Novel adeno-associated viruses from rhesus monkeys as vectors for human gene therapy. Proc. Natl. Acad. Sci. USA 2002, 99, 11854–11859. [Google Scholar] [CrossRef] [Green Version]
  36. Wu, Z.; Miller, E.; Agbandje-McKenna, M.; Samulski, R.J. Alpha2,3 and alpha2,6 N-linked sialic acids facilitate efficient binding and transduction by adeno-associated virus types 1 and 6. J. Virol. 2006, 80, 9093–9103. [Google Scholar] [CrossRef] [Green Version]
  37. Li, C.; Samulski, R.J. Engineering adeno-associated virus vectors for gene therapy. Nat. Rev. Genet. 2020, 21, 255–272. [Google Scholar] [CrossRef]
  38. Rabinowitz, J.E.; Rolling, F.; Li, C.; Conrath, H.; Xiao, W.; Xiao, X.; Samulski, R.J. Cross-packaging of a single adeno-associated virus (AAV) type 2 vector genome into multiple AAV serotypes enables transduction with broad specificity. J. Virol. 2002, 76, 791–801. [Google Scholar] [CrossRef] [Green Version]
  39. Zolotukhin, S.; Potter, M.; Zolotukhin, I.; Sakai, Y.; Loiler, S.; Fraites, T.J., Jr.; Chiodo, V.A.; Phillipsberg, T.; Muzyczka, N.; Hauswirth, W.W.; et al. Production and purification of serotype 1, 2, and 5 recombinant adeno-associated viral vectors. Methods 2002, 28, 158–167. [Google Scholar] [CrossRef]
  40. Robert, M.A.; Chahal, P.S.; Audy, A.; Kamen, A.; Gilbert, R.; Gaillet, B. Manufacturing of recombinant adeno-associated viruses using mammalian expression platforms. Biotechnol. J. 2017, 12, 1600193. [Google Scholar] [CrossRef]
  41. Mary, B.; Maurya, S.; Arumugam, S.; Kumar, V.; Jayandharan, G.R. Post-translational modifications in capsid proteins of recombinant adeno-associated virus (AAV) 1-rh10 serotypes. FEBS J. 2019, 286, 4964–4981. [Google Scholar] [CrossRef]
  42. Huang, L.-Y.; Patel, A.; Ng, R.; Miller, E.B.; Halder, S.; McKenna, R.; Asokan, A.; Agbandje-McKenna, M.; Banks, L. Characterization of the Adeno-Associated Virus 1 and 6 Sialic Acid Binding Site. J. Virol. 2016, 90, 5219–5230. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Chao, H.; Liu, Y.; Rabinowitz, J.; Li, C.; Samulski, R.J.; Walsh, C.E. Several log increase in therapeutic transgene delivery by distinct adeno-associated viral serotype vectors. Mol. Ther. J. Am. Soc. Gene Ther. 2000, 2, 619–623. [Google Scholar] [CrossRef] [PubMed]
  44. Hauck, B.; Xiao, W. Characterization of tissue tropism determinants of adeno-associated virus type 1. J. Virol. 2003, 77, 2768–2774. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Lisowski, L.; Tay, S.S.; Alexander, I.E. Adeno-associated virus serotypes for gene therapeutics. Curr. Opin. Pharmacol. 2015, 24, 59–67. [Google Scholar] [CrossRef]
  46. Lu, Y.; Choi, Y.K.; Campbell-Thompson, M.; Li, C.; Tang, Q.; Crawford, J.M.; Flotte, T.R.; Song, S. Therapeutic level of functional human alpha 1 antitrypsin (hAAT) secreted from murine muscle transduced by adeno-associated virus (rAAV1) vector. J. Gene Med. 2006, 8, 730–735. [Google Scholar] [CrossRef]
  47. Wang, C.; Wang, C.M.; Clark, K.R.; Sferra, T.J. Recombinant AAV serotype 1 transduction efficiency and tropism in the murine brain. Gene Ther. 2003, 10, 1528–1534. [Google Scholar] [CrossRef] [Green Version]
  48. Su, H.; Yeghiazarians, Y.; Lee, A.; Huang, Y.; Arakawa-Hoyt, J.; Ye, J.; Orcino, G.; Grossman, W.; Kan, Y.W. AAV serotype 1 mediates more efficient gene transfer to pig myocardium than AAV serotype 2 and plasmid. J. Gene Med. 2008, 10, 33–41. [Google Scholar] [CrossRef]
  49. Palomeque, J.; Chemaly, E.R.; Colosi, P.; Wellman, J.A.; Zhou, S.; Del Monte, F.; Hajjar, R.J. Efficiency of eight different AAV serotypes in transducing rat myocardium in vivo. Gene Ther. 2007, 14, 989–997. [Google Scholar] [CrossRef] [Green Version]
  50. Chen, S.; Kapturczak, M.; Loiler, S.A.; Zolotukhin, S.; Glushakova, O.Y.; Madsen, K.M.; Samulski, R.J.; Hauswirth, W.W.; Campbell-Thompson, M.; Berns, K.I.; et al. Efficient transduction of vascular endothelial cells with recombinant adeno-associated virus serotype 1 and 5 vectors. Hum. Gene Ther. 2005, 16, 235–247. [Google Scholar] [CrossRef] [Green Version]
  51. Auricchio, A.; Kobinger, G.; Anand, V.; Hildinger, M.; O’Connor, E.; Maguire, A.M.; Wilson, J.M.; Bennett, J. Exchange of surface proteins impacts on viral vector cellular specificity and transduction characteristics: The retina as a model. Hum. Mol. Genet. 2001, 10, 3075–3081. [Google Scholar] [CrossRef] [Green Version]
  52. Srivastava, A. In vivo tissue-tropism of adeno-associated viral vectors. Curr. Opin. Virol. 2016, 21, 75–80. [Google Scholar] [CrossRef] [Green Version]
  53. Summerford, C.; Samulski, R.J. Membrane-Associated Heparan Sulfate Proteoglycan Is a Receptor for Adeno-Associated Virus Type 2 Virions. J. Virol. 1998, 72, 1438–1445. [Google Scholar] [CrossRef] [Green Version]
  54. Opie, S.R.; Warrington, K.H., Jr.; Agbandje-McKenna, M.; Zolotukhin, S.; Muzyczka, N. Identification of amino acid residues in the capsid proteins of adeno-associated virus type 2 that contribute to heparan sulfate proteoglycan binding. J. Virol. 2003, 77, 6995–7006. [Google Scholar] [CrossRef] [Green Version]
  55. Auricchio, A.; Sena-Esteves, M.; Gao, G. Purification of Recombinant Adeno-Associated Virus 2 (rAAV2) by Heparin Column Affinity Chromatography. Cold Spring Harb. Protoc. 2020, 2020, 095620. [Google Scholar] [CrossRef]
  56. Qing, K.; Mah, C.; Hansen, J.; Zhou, S.; Dwarki, V.; Srivastava, A.J. Human fibroblast growth factor receptor 1 is a co-receptor for infection by adeno-associated virus 2. Nat. Med. 1999, 5, 71–77. [Google Scholar] [CrossRef]
  57. Summerford, C.; Bartlett, J.S.; Samulski, R.J.J. αVβ5 integrin: A co-receptor for adeno-associated virus type 2 infection. Nat. Med. 1999, 5, 78–82. [Google Scholar] [CrossRef]
  58. Asokan, A.; Hamra, J.B.; Govindasamy, L.; Agbandje-McKenna, M.; Samulski, R.J.J. Adeno-associated virus type 2 contains an integrin α5β1 binding domain essential for viral cell entry. J. Virol. 2006, 80, 8961–8969. [Google Scholar] [CrossRef] [Green Version]
  59. Kashiwakura, Y.; Tamayose, K.; Iwabuchi, K.; Hirai, Y.; Shimada, T.; Matsumoto, K.; Nakamura, T.; Watanabe, M.; Oshimi, K.; Daida, H.J. Hepatocyte growth factor receptor is a coreceptor for adeno-associated virus type 2 infection. J. Virol. 2005, 79, 609–614. [Google Scholar] [CrossRef] [Green Version]
  60. Akache, B.; Grimm, D.; Pandey, K.; Yant, S.R.; Xu, H.; Kay, M.A.J. The 37/67-kilodalton laminin receptor is a receptor for adeno-associated virus serotypes 8, 2, 3, and 9. J. Virol. 2006, 80, 9831–9836. [Google Scholar] [CrossRef] [Green Version]
  61. Kurzeder, C.; Koppold, B.; Sauer, G.; Pabst, S.; Kreienberg, R.; Deissler, H.J. CD9 promotes adeno-associated virus type 2 infection of mammary carcinoma cells with low cell surface expression of heparan sulphate proteoglycans. Int. J. Mol. Med. 2007, 19, 325–333. [Google Scholar] [CrossRef]
  62. Takeda, S.-i.; Takahashi, M.; Mizukami, H.; Kobayashi, E.; Takeuchi, K.; Hakamata, Y.; Kaneko, T.; Yamamoto, H.; Ito, C.; Ozawa, K.J. Successful gene transfer using adeno-associated virus vectors into the kidney: Comparison among adeno-associated virus serotype 1–5 vectors in vitro and in vivo. Nephron Exp. Nephrol. 2004, 96, e119–e126. [Google Scholar] [CrossRef] [PubMed]
  63. Qi, Y.F.; Li, Q.H.; Shenoy, V.; Zingler, M.; Jun, J.Y.; Verma, A.; Katovich, M.J.; Raizada, M.K. Comparison of the transduction efficiency of tyrosine-mutant adeno-associated virus serotype vectors in kidney. Clin. Exp. Pharmacol. Physiol. 2013, 40, 53–55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Haddad, G.; Kölling, M.; Wegmann, U.A.; Dettling, A.; Seeger, H.; Schmitt, R.; Soerensen-Zender, I.; Haller, H.; Kistler, A.D.; Dueck, A.J. Renal AAV2-mediated overexpression of long non-coding RNA H19 attenuates ischemic acute kidney injury through sponging of microRNA-30a-5p. J. Am. Soc. Nephrol. 2021, 32, 323–341. [Google Scholar] [CrossRef] [PubMed]
  65. Logan, G.J.; Dane, A.P.; Hallwirth, C.V.; Smyth, C.M.; Wilkie, E.E.; Amaya, A.K.; Zhu, E.; Khandekar, N.; Ginn, S.L.; Liao, S.H. Identification of liver-specific enhancer–promoter activity in the 3′ untranslated region of the wild-type AAV2 genome. Nat. Genet. 2017, 49, 1267–1273. [Google Scholar] [CrossRef] [PubMed]
  66. Niemeyer, G.P.; Herzog, R.W.; Mount, J.; Arruda, V.R.; Tillson, D.M.; Hathcock, J.; van Ginkel, F.W.; High, K.A.; Lothrop, C.D., Jr. The Journal of the American Society of Hematology. Long-term correction of inhibitor-prone hemophilia B dogs treated with liver-directed AAV2-mediated factor IX gene therapy. Blood J. Am. Soc. Hematol. 2009, 113, 797–806. [Google Scholar]
  67. Petrs-Silva, H.; Dinculescu, A.; Li, Q.; Deng, W.-T.; Pang, J.-j.; Min, S.-H.; Chiodo, V.; Neeley, A.W.; Govindasamy, L.; Bennett, A.; et al. Novel Properties of Tyrosine-mutant AAV2 Vectors in the Mouse Retina. Mol. Ther. 2011, 19, 293–301. [Google Scholar] [CrossRef]
  68. Yin, L.; Greenberg, K.; Hunter, J.J.; Dalkara, D.; Kolstad, K.D.; Masella, B.D.; Wolfe, R.; Visel, M.; Stone, D.; Libby, R.T.; et al. Intravitreal injection of AAV2 transduces macaque inner retina. Investig. Ophthalmol. Vis. Sci. 2011, 52, 2775–2783. [Google Scholar] [CrossRef]
  69. Wassmer, S.J.; Carvalho, L.S.; György, B.; Vandenberghe, L.H.; Maguire, C.A. Exosome-associated AAV2 vector mediates robust gene delivery into the murine retina upon intravitreal injection. J. Sci. Rep. 2017, 7, 45329. [Google Scholar] [CrossRef] [Green Version]
  70. Griffey, M.; Bible, E.; Vogler, C.; Levy, B.; Gupta, P.; Cooper, J.; Sands, M.S. Adeno-associated virus 2-mediated gene therapy decreases autofluorescent storage material and increases brain mass in a murine model of infantile neuronal ceroid lipofuscinosis. Neurobiol. Dis. 2004, 16, 360–369. [Google Scholar] [CrossRef]
  71. Griffey, M.A.; Wozniak, D.; Wong, M.; Bible, E.; Johnson, K.; Rothman, S.M.; Wentz, A.E.; Cooper, J.D.; Sands, M.S. CNS-directed AAV2-mediated gene therapy ameliorates functional deficits in a murine model of infantile neuronal ceroid lipofuscinosis. Mol. Ther. 2006, 13, 538–547. [Google Scholar] [CrossRef]
  72. Pankajakshan, D.; Makinde, T.O.; Gaurav, R.; Del Core, M.; Hatzoudis, G.; Pipinos, I.; Agrawal, D.K. Successful Transfection of Genes Using AAV-2/9 Vector in Swine Coronary and Peripheral Arteries. J. Surg. Res. 2012, 175, 169–175. [Google Scholar] [CrossRef] [Green Version]
  73. Rabinowitz, J.E.; Bowles, D.E.; Faust, S.M.; Ledford, J.G.; Cunningham, S.E.; Samulski, R.J. Cross-Dressing the Virion: The Transcapsidation of Adeno-Associated Virus Serotypes Functionally Defines Subgroups. J. Virol. 2004, 78, 4421–4432. [Google Scholar] [CrossRef] [Green Version]
  74. Blackburn, S.D.; Steadman, R.A.; Johnson, F.B. Attachment of adeno-associated virus type 3H to fibroblast growth factor receptor 1. Arch. Virol. 2006, 151, 617–623. [Google Scholar] [CrossRef]
  75. Ling, C.; Wang, Y.; Zhang, Y.; Ejjigani, A.; Yin, Z.; Lu, Y.; Wang, L.; Wang, M.; Li, J.; Hu, Z.; et al. Selective in vivo targeting of human liver tumors by optimized AAV3 vectors in a murine xenograft model. Hum. Gene Ther. 2014, 25, 1023–1034. [Google Scholar] [CrossRef] [Green Version]
  76. Ling, C.; Yin, Z.; Li, J.; Zhang, D.; Aslanidi, G.; Srivastava, A. Strategies to generate high-titer, high-potency recombinant AAV3 serotype vectors. Mol. Ther. Methods Clin. Dev. 2016, 3, 16029. [Google Scholar] [CrossRef]
  77. Handa, A.; Muramatsu, S.I.; Qiu, J.; Mizukami, H.; Brown, K.E. Adeno-associated virus (AAV)-3-based vectors transduce haematopoietic cells not susceptible to transduction with AAV-2-based vectors. J. Gen. Virol. 2000, 81, 2077–2084. [Google Scholar] [CrossRef]
  78. Cheng, B.; Ling, C.; Dai, Y.; Lu, Y.; Glushakova, L.G.; Gee, S.W.Y.; McGoogan, K.E.; Aslanidi, G.V.; Park, M.; Stacpoole, P.W.; et al. Development of optimized AAV3 serotype vectors: Mechanism of high-efficiency transduction of human liver cancer cells. Gene Ther. 2012, 19, 375–384. [Google Scholar] [CrossRef] [Green Version]
  79. Muramatsu, S.; Mizukami, H.; Young, N.S.; Brown, K.E. Nucleotide sequencing and generation of an infectious clone of adeno-associated virus 3. Virology 1996, 221, 208–217. [Google Scholar] [CrossRef] [Green Version]
  80. Li, S.; Ling, C.; Zhong, L.; Li, M.; Su, Q.; He, R.; Tang, Q.; Greiner, D.L.; Shultz, L.D.; Brehm, M.A.; et al. Efficient and Targeted Transduction of Nonhuman Primate Liver With Systemically Delivered Optimized AAV3B Vectors. Mol. Ther. 2015, 23, 1867–1876. [Google Scholar] [CrossRef] [Green Version]
  81. Vercauteren, K.; Hoffman, B.E.; Zolotukhin, I.; Keeler, G.D.; Xiao, J.W.; Basner-Tschakarjan, E.; High, K.A.; Ertl, H.C.J.; Rice, C.M.; Srivastava, A.; et al. Superior In vivo Transduction of Human Hepatocytes Using Engineered AAV3 Capsid. Mol. Ther. 2016, 24, 1042–1049. [Google Scholar] [CrossRef] [Green Version]
  82. Liu, Y.; Okada, T.; Sheykholeslami, K.; Shimazaki, K.; Nomoto, T.; Muramatsu, S.; Kanazawa, T.; Takeuchi, K.; Ajalli, R.; Mizukami, H.; et al. Specific and efficient transduction of Cochlear inner hair cells with recombinant adeno-associated virus type 3 vector. Mol. Ther. J. Am. Soc. Gene Ther. 2005, 12, 725–733. [Google Scholar] [CrossRef] [PubMed]
  83. Padron, E.; Bowman, V.; Kaludov, N.; Govindasamy, L.; Levy, H.; Nick, P.; McKenna, R.; Muzyczka, N.; Chiorini, J.A.; Baker, T.S.; et al. Structure of adeno-associated virus type 4. J. Virol. 2005, 79, 5047–5058. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Parks, W.P.; Boucher, D.W.; Melnick, J.L.; Taber, L.H.; Yow, M.D. Seroepidemiological and ecological studies of the adenovirus-associated satellite viruses. Infect. Immun. 1970, 2, 716–722. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Blacklow, N.R.; Hoggan, M.D.; Rowe, W.P. Serologic evidence for human infection with adenovirus-associated viruses. J. Natl. Cancer Inst. 1968, 40, 319–327. [Google Scholar] [PubMed]
  86. Dreizin, R.S.; Zhuravel, T.F.; Tarasova, A.B.; Sobolev, S.G.; Kozlov, V.G. Experimental infection of green monkeys with adenoassociated virus. Vopr. Virusol. 1981, 1, 82–89. [Google Scholar]
  87. Chiorini, J.A.; Yang, L.; Liu, Y.; Safer, B.; Kotin, R.M. Cloning of adeno-associated virus type 4 (AAV4) and generation of recombinant AAV4 particles. J. Virol. 1997, 71, 6823–6833. [Google Scholar] [CrossRef] [Green Version]
  88. Kaludov, N.; Brown, K.E.; Walters, R.W.; Zabner, J.; Chiorini, J.A. Adeno-associated virus serotype 4 (AAV4) and AAV5 both require sialic acid binding for hemagglutination and efficient transduction but differ in sialic acid linkage specificity. J. Virol. 2001, 75, 6884–6893. [Google Scholar] [CrossRef] [Green Version]
  89. Kaludov, N.; Handelman, B.; Chiorini, J.A. Scalable purification of adeno-associated virus type 2, 4, or 5 using ion-exchange chromatography. Hum. Gene Ther. 2002, 13, 1235–1243. [Google Scholar] [CrossRef]
  90. Davidson, B.L.; Stein, C.S.; Heth, J.A.; Martins, I.; Kotin, R.M.; Derksen, T.A.; Zabner, J.; Ghodsi, A.; Chiorini, J.A. 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 2000, 97, 3428–3432. [Google Scholar] [CrossRef]
  91. Weber, M.; Rabinowitz, J.; Provost, N.; Conrath, H.; Folliot, S.; Briot, D.; Chérel, Y.; Chenuaud, P.; Samulski, J.; Moullier, P.; et al. Recombinant adeno-associated virus serotype 4 mediates unique and exclusive long-term transduction of retinal pigmented epithelium in rat, dog, and nonhuman primate after subretinal delivery. Mol. Ther. J. Am. Soc. Gene Ther. 2003, 7, 774–781. [Google Scholar] [CrossRef]
  92. Zincarelli, C.; Soltys, S.; Rengo, G.; Rabinowitz, J.E. Analysis of AAV serotypes 1–9 mediated gene expression and tropism in mice after systemic injection. Mol. Ther. 2008, 16, 1073–1080. [Google Scholar] [CrossRef]
  93. Bantel-Schaal, U.; zur Hausen, H. Characterization of the DNA of a defective human parvovirus isolated from a genital site. Virology 1984, 134, 52–63. [Google Scholar] [CrossRef]
  94. Stutika, C.; Hüser, D.; Weger, S.; Rutz, N.; Heßler, M.; Heilbronn, R. Definition of herpes simplex virus helper functions for the replication of adeno-associated virus type 5. J. Gen. Virol. 2015, 96, 840–850. [Google Scholar] [CrossRef]
  95. Bantel-Schaal, U.; Delius, H.; Schmidt, R.; Hausen, H.Z. Human Adeno-Associated Virus Type 5 Is Only Distantly Related to Other Known Primate Helper-Dependent Parvoviruses. J. Virol. 1999, 73, 939–947. [Google Scholar] [CrossRef] [Green Version]
  96. Chiorini, J.A.; Kim, F.; Yang, L.; Kotin, R.M. Cloning and characterization of adeno-associated virus type 5. J. Virol. 1999, 73, 1309–1319. [Google Scholar] [CrossRef] [Green Version]
  97. Bantel-Schaal, U.; Hub, B.; Kartenbeck, J. Endocytosis of adeno-associated virus type 5 leads to accumulation of virus particles in the Golgi compartment. J. Virol. 2002, 76, 2340–2349. [Google Scholar] [CrossRef] [Green Version]
  98. Zabner, J.; Seiler, M.; Walters, R.; Kotin, R.M.; Fulgeras, W.; Davidson, B.L.; Chiorini, J.A. Adeno-associated virus type 5 (AAV5) but not AAV2 binds to the apical surfaces of airway epithelia and facilitates gene transfer. J. Virol. 2000, 74, 3852–3858. [Google Scholar] [CrossRef] [Green Version]
  99. Grimm, D.; Kay, M.A. From virus evolution to vector revolution: Use of naturally occurring serotypes of adeno-associated virus (AAV) as novel vectors for human gene therapy. Curr. Gene Ther. 2003, 3, 281–304. [Google Scholar] [CrossRef]
  100. Walters, R.W.; Yi, S.M.; Keshavjee, S.; Brown, K.E.; Welsh, M.J.; Chiorini, J.A.; Zabner, J. Binding of adeno-associated virus type 5 to 2,3-linked sialic acid is required for gene transfer. J. Biol. Chem. 2001, 276, 20610–20616. [Google Scholar] [CrossRef] [Green Version]
  101. Henckaerts, E.; Linden, R.M. Adeno-associated virus: A key to the human genome? Future Virol. 2010, 5, 555–574. [Google Scholar] [CrossRef] [Green Version]
  102. Di Pasquale, G.; Davidson, B.L.; Stein, C.S.; Martins, I.; Scudiero, D.; Monks, A.; Chiorini, J.A. Identification of PDGFR as a receptor for AAV-5 transduction. Nat. Med. 2003, 9, 1306–1312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Nonnenmacher, M.; Weber, T. Intracellular transport of recombinant adeno-associated virus vectors. Gene Ther. 2012, 19, 649–658. [Google Scholar] [CrossRef] [PubMed]
  104. Auricchio, A.; O’Connor, E.; Hildinger, M.; Wilson, J.M. A Single-Step Affinity Column for Purification of Serotype-5 Based Adeno-associated Viral Vectors. Mol. Ther. 2001, 4, 372–374. [Google Scholar] [CrossRef] [PubMed]
  105. Walters, R.W.; Pilewski, J.M.; Chiorini, J.A.; Zabner, J. Secreted and Transmembrane Mucins Inhibit Gene Transfer with AAV4 More Efficiently than AAV5*. J. Biol. Chem. 2002, 277, 23709–23713. [Google Scholar] [CrossRef] [Green Version]
  106. Yang, G.S.; Schmidt, M.; Yan, Z.; Lindbloom, J.D.; Harding, T.C.; Donahue, B.A.; Engelhardt, J.F.; Kotin, R.; Davidson, B.L. Virus-mediated transduction of murine retina with adeno-associated virus: Effects of viral capsid and genome size. J. Virol. 2002, 76, 7651–7660. [Google Scholar] [CrossRef] [Green Version]
  107. Alisky, J.M.; Hughes, S.M.; Sauter, S.L.; Jolly, D.; Dubensky, T.W., Jr.; Staber, P.D.; Chiorini, J.A.; Davidson, B.L. Transduction of murine cerebellar neurons with recombinant FIV and AAV5 vectors. Neuroreport 2000, 11, 2669–2673. [Google Scholar] [CrossRef]
  108. Burger, C.; Gorbatyuk, O.S.; Velardo, M.J.; Peden, C.S.; Williams, P.; Zolotukhin, S.; Reier, P.J.; Mandel, R.J.; Muzyczka, N. Recombinant AAV viral vectors pseudotyped with viral capsids from serotypes 1, 2, and 5 display differential efficiency and cell tropism after delivery to different regions of the central nervous system. Mol. Ther. J. Am. Soc. Gene Ther. 2004, 10, 302–317. [Google Scholar] [CrossRef]
  109. Ding, W.; Yan, Z.; Zak, R.; Saavedra, M.; Rodman, D.M.; Engelhardt, J.F. Second-strand genome conversion of adeno-associated virus type 2 (AAV-2) and AAV-5 is not rate limiting following apical infection of polarized human airway epithelia. J. Virol. 2003, 77, 7361–7366. [Google Scholar] [CrossRef] [Green Version]
  110. Auricchio, A.; O’Connor, E.; Weiner, D.; Gao, G.P.; Hildinger, M.; Wang, L.; Calcedo, R.; Wilson, J.M. Noninvasive gene transfer to the lung for systemic delivery of therapeutic proteins. J. Clin. Investig. 2002, 110, 499–504. [Google Scholar] [CrossRef]
  111. Seiler, M.P.; Miller, A.D.; Zabner, J.; Halbert, C.L. Adeno-associated virus types 5 and 6 use distinct receptors for cell entry. Hum. Gene Ther. 2006, 17, 10–19. [Google Scholar] [CrossRef]
  112. Pañeda, A.; Vanrell, L.; Mauleon, I.; Crettaz, J.S.; Berraondo, P.; Timmermans, E.J.; Beattie, S.G.; Twisk, J.; van Deventer, S.; Prieto, J.; et al. Effect of adeno-associated virus serotype and genomic structure on liver transduction and biodistribution in mice of both genders. Hum. Gene Ther. 2009, 20, 908–917. [Google Scholar] [CrossRef]
  113. Xiao, W.; Chirmule, N.; Berta, S.C.; McCullough, B.; Gao, G.; Wilson, J.M. Gene Therapy Vectors Based on Adeno-Associated Virus Type 1. J. Virol. 1999, 73, 3994–4003. [Google Scholar] [CrossRef] [Green Version]
  114. Rutledge, E.A.; Halbert, C.L.; Russell, D.W. Infectious Clones and Vectors Derived from Adeno-Associated Virus (AAV) Serotypes Other Than AAV Type 2. J. Virol. 1998, 72, 309–319. [Google Scholar] [CrossRef] [Green Version]
  115. Korneyenkov, M.A.; Zamyatnin, A.A., Jr. Next Step in Gene Delivery: Modern Approaches and Further Perspectives of AAV Tropism Modification. Pharmaceutics 2021, 13, 750. [Google Scholar] [CrossRef]
  116. Ng, R.; Govindasamy, L.; Gurda, B.L.; McKenna, R.; Kozyreva, O.G.; Samulski, R.J.; Parent, K.N.; Baker, T.S.; Agbandje-McKenna, M. Structural characterization of the dual glycan binding adeno-associated virus serotype 6. J. Virol. 2010, 84, 12945–12957. [Google Scholar] [CrossRef] [Green Version]
  117. Wu, Z.; Asokan, A.; Grieger, J.C.; Govindasamy, L.; Agbandje-McKenna, M.; Samulski, R.J. Single amino acid changes can influence titer, heparin binding, and tissue tropism in different adeno-associated virus serotypes. J. Virol. 2006, 80, 11393–11397. [Google Scholar] [CrossRef] [Green Version]
  118. Weller, M.L.; Amornphimoltham, P.; Schmidt, M.; Wilson, P.A.; Gutkind, J.S.; Chiorini, J.A. Epidermal growth factor receptor is a co-receptor for adeno-associated virus serotype 6. Nat. Med. 2010, 16, 662–664. [Google Scholar] [CrossRef] [Green Version]
  119. Halbert, C.L.; Allen, J.M.; Miller, A.D. Adeno-associated virus type 6 (AAV6) vectors mediate efficient transduction of airway epithelial cells in mouse lungs compared to that of AAV2 vectors. J. Virol. 2001, 75, 6615–6624. [Google Scholar] [CrossRef] [Green Version]
  120. Halbert, C.L.; Madtes, D.K.; Vaughan, A.E.; Wang, Z.; Storb, R.; Tapscott, S.J.; Miller, A.D. Expression of human alpha1-antitrypsin in mice and dogs following AAV6 vector-mediated gene transfer to the lungs. Mol. Ther. J. Am. Soc. Gene Ther. 2010, 18, 1165–1172. [Google Scholar] [CrossRef]
  121. Grimm, D.; Zhou, S.; Nakai, H.; Thomas, C.E.; Storm, T.A.; Fuess, S.; Matsushita, T.; Allen, J.; Surosky, R.; Lochrie, M.; et al. Preclinical in vivo evaluation of pseudotyped adeno-associated virus vectors for liver gene therapy. Blood 2003, 102, 2412–2419. [Google Scholar] [CrossRef]
  122. Blankinship, M.J.; Gregorevic, P.; Allen, J.M.; Harper, S.Q.; Harper, H.; Halbert, C.L.; Miller, A.D.; Chamberlain, J.S. Efficient transduction of skeletal muscle using vectors based on adeno-associated virus serotype 6. Mol. Ther. J. Am. Soc. Gene Ther. 2004, 10, 671–678. [Google Scholar] [CrossRef]
  123. Wang, Z.; Kuhr, C.S.; Allen, J.M.; Blankinship, M.; Gregorevic, P.; Chamberlain, J.S.; Tapscott, S.J.; Storb, R. Sustained AAV-mediated dystrophin expression in a canine model of Duchenne muscular dystrophy with a brief course of immunosuppression. Mol. Ther. J. Am. Soc. Gene Ther. 2007, 15, 1160–1166. [Google Scholar] [CrossRef] [PubMed]
  124. Qiao, C.; Zhang, W.; Yuan, Z.; Shin, J.H.; Li, J.; Jayandharan, G.R.; Zhong, L.; Srivastava, A.; Xiao, X.; Duan, D. Adeno-associated virus serotype 6 capsid tyrosine-to-phenylalanine mutations improve gene transfer to skeletal muscle. Hum. Gene Ther. 2010, 21, 1343–1348. [Google Scholar] [CrossRef] [PubMed]
  125. Zincarelli, C.; Soltys, S.; Rengo, G.; Koch, W.J.; Rabinowitz, J.E. Comparative cardiac gene delivery of adeno-associated virus serotypes 1–9 reveals that AAV6 mediates the most efficient transduction in mouse heart. Clin. Transl. Sci. 2010, 3, 81–89. [Google Scholar] [CrossRef] [PubMed]
  126. Raake, P.W.; Hinkel, R.; Müller, S.; Delker, S.; Kreuzpointner, R.; Kupatt, C.; Katus, H.A.; Kleinschmidt, J.A.; Boekstegers, P.; Müller, O.J. Cardio-specific long-term gene expression in a porcine model after selective pressure-regulated retroinfusion of adeno-associated viral (AAV) vectors. Gene Ther. 2008, 15, 12–17. [Google Scholar] [CrossRef] [Green Version]
  127. Bish, L.T.; Sleeper, M.M.; Reynolds, C.; Gazzara, J.; Withnall, E.; Singletary, G.E.; Buchlis, G.; Hui, D.; High, K.A.; Gao, G.; et al. Cardiac gene transfer of short hairpin RNA directed against phospholamban effectively knocks down gene expression but causes cellular toxicity in canines. Hum. Gene Ther. 2011, 22, 969–977. [Google Scholar] [CrossRef] [Green Version]
  128. White, J.D.; Thesier, D.M.; Swain, J.B.; Katz, M.G.; Tomasulo, C.; Henderson, A.; Wang, L.; Yarnall, C.; Fargnoli, A.; Sumaroka, M.; et al. Myocardial gene delivery using molecular cardiac surgery with recombinant adeno-associated virus vectors in vivo. Gene Ther. 2011, 18, 546–552. [Google Scholar] [CrossRef]
  129. Gao, G.; Vandenberghe, L.H.; Alvira, M.R.; Lu, Y.; Calcedo, R.; Zhou, X.; Wilson, J.M. Clades of Adeno-associated viruses are widely disseminated in human tissues. J. Virol. 2004, 78, 6381–6388. [Google Scholar] [CrossRef] [Green Version]
  130. Denby, L.; Nicklin, S.A.; Baker, A.H. Adeno-associated virus (AAV)-7 and -8 poorly transduce vascular endothelial cells and are sensitive to proteasomal degradation. Gene Ther. 2005, 12, 1534–1538. [Google Scholar] [CrossRef]
  131. Mietzsch, M.; Broecker, F.; Reinhardt, A.; Seeberger, P.H.; Heilbronn, R. Differential Adeno-Associated Virus Serotype-Specific Interaction Patterns with Synthetic Heparins and Other Glycans. J. Virol. 2014, 88, 2991–3003. [Google Scholar] [CrossRef] [Green Version]
  132. Calcedo, R.; Vandenberghe, L.H.; Gao, G.; Lin, J.; Wilson, J.M. Worldwide Epidemiology of Neutralizing Antibodies to Adeno-Associated Viruses. J. Infect. Dis. 2009, 199, 381–390. [Google Scholar] [CrossRef]
  133. Louboutin, J.-P.; Wang, L.; Wilson, J.M. Gene transfer into skeletal muscle using novel AAV serotypes. J. Gene Med. A Cross-Discip. J. Res. Sci. Gene Transf. Clin. Appl. 2005, 7, 442–451. [Google Scholar] [CrossRef]
  134. Samaranch, L.; Salegio, E.A.; San Sebastian, W.; Kells, A.P.; Bringas, J.R.; Forsayeth, J.; Bankiewicz, K.S. Strong cortical and spinal cord transduction after AAV7 and AAV9 delivery into the cerebrospinal fluid of nonhuman primates. Hum. Gene Ther. 2013, 24, 526–532. [Google Scholar] [CrossRef] [Green Version]
  135. Taymans, J.M.; Vandenberghe, L.H.; Haute, C.V.; Thiry, I.; Deroose, C.M.; Mortelmans, L.; Wilson, J.M.; Debyser, Z.; Baekelandt, V. Comparative analysis of adeno-associated viral vector serotypes 1, 2, 5, 7, and 8 in mouse brain. Hum. Gene Ther. 2007, 18, 195–206. [Google Scholar] [CrossRef]
  136. Allocca, M.; Mussolino, C.; Garcia-Hoyos, M.; Sanges, D.; Iodice, C.; Petrillo, M.; Vandenberghe, L.H.; Wilson, J.M.; Marigo, V.; Surace, E.M.; et al. Novel adeno-associated virus serotypes efficiently transduce murine photoreceptors. J. Virol. 2007, 81, 11372–11380. [Google Scholar] [CrossRef] [Green Version]
  137. Okada, T.; Nonaka-Sarukawa, M.; Uchibori, R.; Kinoshita, K.; Hayashita-Kinoh, H.; Nitahara-Kasahara, Y.; Takeda, S.i.; Ozawa, K. Scalable purification of adeno-associated virus serotype 1 (AAV1) and AAV8 vectors, using dual ion-exchange adsorptive membranes. Hum. Gene Ther. 2009, 20, 1013–1021. [Google Scholar] [CrossRef]
  138. Davidoff, A.M.; Ng, C.Y.; Sleep, S.; Gray, J.; Azam, S.; Zhao, Y.; McIntosh, J.H.; Karimipoor, M.; Nathwani, A.C. Purification of recombinant adeno-associated virus type 8 vectors by ion exchange chromatography generates clinical grade vector stock. J. Virol. Methods 2004, 121, 209–215. [Google Scholar] [CrossRef]
  139. Lane, M.D.; Nam, H.-J.; Padron, E.; Gurda-Whitaker, B.; Kohlbrenner, E.; Aslanidi, G.; Byrne, B.; McKenna, R.; Muzyczka, N.; Zolotukhin, S.; et al. Production, purification, crystallization and preliminary X-ray analysis of adeno-associated virus serotype 8. Acta Crystallogr. Sect. F Struct. Biol. Cryst. Commun. 2005, 61, 558–561. [Google Scholar] [CrossRef] [Green Version]
  140. Miyake, K.; Miyake, N.; Yamazaki, Y.; Shimada, T.; Hirai, Y.J. Serotype-independent method of recombinant adeno-associated virus (AAV) vector production and purification. J. Nippon. Med. Sch. 2012, 79, 394–402. [Google Scholar] [CrossRef] [Green Version]
  141. Nam, H.-J.; Lane, M.D.; Padron, E.; Gurda, B.; McKenna, R.; Kohlbrenner, E.; Aslanidi, G.; Byrne, B.; Muzyczka, N.; Zolotukhin, S.; et al. Structure of Adeno-Associated Virus Serotype 8, a Gene Therapy Vector. J. Virol. 2007, 81, 12260–12271. [Google Scholar] [CrossRef] [Green Version]
  142. Cabanes-Creus, M.; Navarro, R.G.; Zhu, E.; Baltazar, G.; Liao, S.H.Y.; Drouyer, M.; Amaya, A.K.; Scott, S.; Nguyen, L.H.; Westhaus, A.; et al. Novel human liver-tropic AAV variants define transferable domains that markedly enhance the human tropism of AAV7 and AAV8. Mol. Ther. Methods Clin. Dev. 2022, 24, 88–101. [Google Scholar] [CrossRef] [PubMed]
  143. Nakai, H.; Fuess, S.; Storm, T.A.; Muramatsu, S.; Nara, Y.; Kay, M.A. Unrestricted hepatocyte transduction with adeno-associated virus serotype 8 vectors in mice. J. Virol. 2005, 79, 214–224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Thomas, C.E.; Storm, T.A.; Huang, Z.; Kay, M.A. Rapid uncoating of vector genomes is the key to efficient liver transduction with pseudotyped adeno-associated virus vectors. J. Virol. 2004, 78, 3110–3122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Nam, H.-J.; Gurda, B.L.; McKenna, R.; Potter, M.; Byrne, B.; Salganik, M.; Muzyczka, N.; Agbandje-McKenna, M. Structural Studies of Adeno-Associated Virus Serotype 8 Capsid Transitions Associated with Endosomal Trafficking. J. Virol. 2011, 85, 11791–11799. [Google Scholar] [CrossRef] [Green Version]
  146. Monahan, P.E.; Lothrop, C.D.; Sun, J.; Hirsch, M.L.; Kafri, T.; Kantor, B.; Sarkar, R.; Tillson, D.M.; Elia, J.R.; Samulski, R.J. Proteasome inhibitors enhance gene delivery by AAV virus vectors expressing large genomes in hemophilia mouse and dog models: A strategy for broad clinical application. Mol. Ther. J. Am. Soc. Gene Ther. 2010, 18, 1907–1916. [Google Scholar] [CrossRef]
  147. Jiang, H.; Couto, L.B.; Patarroyo-White, S.; Liu, T.; Nagy, D.; Vargas, J.A.; Zhou, S.; Scallan, C.D.; Sommer, J.; Vijay, S.; et al. Effects of transient immunosuppression on adenoassociated, virus-mediated, liver-directed gene transfer in rhesus macaques and implications for human gene therapy. Blood 2006, 108, 3321–3328. [Google Scholar] [CrossRef] [Green Version]
  148. Wang, Z.; Zhu, T.; Qiao, C.; Zhou, L.; Wang, B.; Zhang, J.; Chen, C.; Li, J.; Xiao, X. Adeno-associated virus serotype 8 efficiently delivers genes to muscle and heart. Nat. Biotechnol. 2005, 23, 321–328. [Google Scholar] [CrossRef]
  149. Wang, A.Y.; Peng, P.D.; Ehrhardt, A.; Storm, T.A.; Kay, M.A. Comparison of adenoviral and adeno-associated viral vectors for pancreatic gene delivery in vivo. Hum. Gene Ther. 2004, 15, 405–413. [Google Scholar] [CrossRef] [Green Version]
  150. Loiler, S.A.; Tang, Q.; Clarke, T.; Campbell-Thompson, M.L.; Chiodo, V.; Hauswirth, W.; Cruz, P.; Perret-Gentil, M.; Atkinson, M.A.; Ramiya, V.K.; et al. Localized gene expression following administration of adeno-associated viral vectors via pancreatic ducts. Mol. Ther. J. Am. Soc. Gene Ther. 2005, 12, 519–527. [Google Scholar] [CrossRef]
  151. Rubin, J.D.; Nguyen, T.V.; Allen, K.L.; Ayasoufi, K.; Barry, M.A. Comparison of gene delivery to the kidney by adenovirus, adeno-associated virus, and lentiviral vectors after intravenous and direct kidney injections. Hum. Gene Ther. 2019, 30, 1559–1571. [Google Scholar] [CrossRef]
  152. Giove, T.J.; Khankhel, Z.S.; Sena-Esteves, M.; Eldred, W.D. Cellular Tropism of Aav8, Aav9 and Aav10 in Mouse Retina. Investig. Ophthalmol. Vis. Sci. 2009, 50, 3013. [Google Scholar]
  153. Ding, K.; Shen, J.; Hafiz, Z.; Hackett, S.F.; Silva, R.L.E.; Khan, M.; Lorenc, V.E.; Chen, D.; Chadha, R.; Zhang, M.; et al. AAV8-vectored suprachoroidal gene transfer produces widespread ocular transgene expression. J. Clin. Investig. 2019, 129, 4901–4911. [Google Scholar] [CrossRef]
  154. Igarashi, T.; Miyake, K.; Asakawa, N.; Miyake, N.; Shimada, T.; Takahashi, H. Direct Comparison of Administration Routes for AAV8-mediated Ocular Gene Therapy. Curr. Eye Res. 2013, 38, 569–577. [Google Scholar] [CrossRef]
  155. Gao, G.; Vandenberghe, L.H.; Wilson, J.M. New recombinant serotypes of AAV vectors. Curr. Gene Ther. 2005, 5, 285–297. [Google Scholar] [CrossRef]
  156. Shen, S.; Bryant, K.D.; Brown, S.M.; Randell, S.H.; Asokan, A. Terminal N-Linked Galactose Is the Primary Receptor for Adeno-associated Virus 9*. J. Biol. Chem. 2011, 286, 13532–13540. [Google Scholar] [CrossRef] [Green Version]
  157. Bell, C.L.; Gurda, B.L.; Van Vliet, K.; Agbandje-McKenna, M.; Wilson, J.M. Identification of the galactose binding domain of the adeno- associated virus serotype 9 capsid. J. Virol. 2012, 86, 7326–7333. [Google Scholar] [CrossRef] [Green Version]
  158. Mitchell, M.; Nam, H.-J.; Carter, A.; McCall, A.; Rence, C.; Bennett, A.; Gurda, B.; McKenna, R.; Porter, M.; Sakai, Y.; et al. Production, purification and preliminary X-ray crystallographic studies of adeno-associated virus serotype 9. Acta Crystallogr. Sect. F Struct. Biol. Cryst. Commun. 2009, 65, 715–718. [Google Scholar] [CrossRef] [Green Version]
  159. Foust, K.D.; Nurre, E.; Montgomery, C.L.; Hernandez, A.; Chan, C.M.; Kaspar, B.K. Intravascular AAV9 preferentially targets neonatal neurons and adult astrocytes. Nat. Biotechnol. 2009, 27, 59–65. [Google Scholar] [CrossRef] [Green Version]
  160. Gray, S.J.; Matagne, V.; Bachaboina, L.; Yadav, S.; Ojeda, S.R.; Samulski, R.J. Preclinical Differences of Intravascular AAV9 Delivery to Neurons and Glia: A Comparative Study of Adult Mice and Nonhuman Primates. Mol. Ther. 2011, 19, 1058–1069. [Google Scholar] [CrossRef]
  161. Bevan, A.K.; Duque, S.; Foust, K.D.; Morales, P.R.; Braun, L.; Schmelzer, L.; Chan, C.M.; McCrate, M.; Chicoine, L.G.; Coley, B.D.; et al. Systemic Gene Delivery in Large Species for Targeting Spinal Cord, Brain, and Peripheral Tissues for Pediatric Disorders. Mol. Ther. 2011, 19, 1971–1980. [Google Scholar] [CrossRef] [Green Version]
  162. Duque, S.; Joussemet, B.; Riviere, C.; Marais, T.; Dubreil, L.; Douar, A.M.; Fyfe, J.; Moullier, P.; Colle, M.A.; Barkats, M. Intravenous administration of self-complementary AAV9 enables transgene delivery to adult motor neurons. Mol. Ther. J. Am. Soc. Gene Ther. 2009, 17, 1187–1196. [Google Scholar] [CrossRef]
  163. Vandenberghe, L.H.; Bell, P.; Maguire, A.M.; Xiao, R.; Hopkins, T.B.; Grant, R.; Bennett, J.; Wilson, J.M.; Xiao, X. AAV9 targets cone photoreceptors in the nonhuman primate retina. PLoS ONE 2013, 8, e53463. [Google Scholar] [CrossRef]
  164. Pacak, C.A.; Mah, C.S.; Thattaliyath, B.D.; Conlon, T.J.; Lewis, M.A.; Cloutier, D.E.; Zolotukhin, I.; Tarantal, A.F.; Byrne, B.J. Recombinant adeno-associated virus serotype 9 leads to preferential cardiac transduction in vivo. Circ. Res. 2006, 99, e3–e9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Asokan, A.; Schaffer, D.V.; Samulski, R.J. The AAV vector toolkit: Poised at the clinical crossroads. Mol. Ther. 2012, 20, 699–708. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Pleger, S.T.; Shan, C.; Ksienzyk, J.; Bekeredjian, R.; Boekstegers, P.; Hinkel, R.; Schinkel, S.; Leuchs, B.; Ludwig, J.; Qiu, G. Cardiac AAV9-S100A1 gene therapy rescues post-ischemic heart failure in a preclinical large animal model. Sci. Transl. Med. 2011, 3, 92ra64. [Google Scholar] [CrossRef] [Green Version]
  167. Bish, L.T.; Morine, K.; Sleeper, M.M.; Sanmiguel, J.; Wu, D.; Gao, G.; Wilson, J.M.; Sweeney, H.L. Adeno-associated virus (AAV) serotype 9 provides global cardiac gene transfer superior to AAV1, AAV6, AAV7, and AAV8 in the mouse and rat. Hum. Gene Ther. 2008, 19, 1359–1368. [Google Scholar] [CrossRef]
  168. Vandendriessche, T.; Thorrez, L.; Acosta-Sanchez, A.; Petrus, I.; Wang, L.; Ma, L.; De Waele, L.; Iwasaki, Y.; Gillijns, V.; Wilson, J.M.; et al. Efficacy and safety of adeno-associated viral vectors based on serotype 8 and 9 vs. lentiviral vectors for hemophilia B gene therapy. J. Thromb. Haemost. JTH 2007, 5, 16–24. [Google Scholar] [CrossRef] [Green Version]
  169. Inagaki, K.; Fuess, S.; Storm, T.A.; Gibson, G.A.; McTiernan, C.F.; Kay, M.A.; Nakai, H. Robust systemic transduction with AAV9 vectors in mice: Efficient global cardiac gene transfer superior to that of AAV8. Mol. Ther. J. Am. Soc. Gene Ther. 2006, 14, 45–53. [Google Scholar] [CrossRef]
  170. Petrs-Silva, H.; Dinculescu, A.; Li, Q.; Min, S.-H.; Chiodo, V.; Pang, J.-J.; Zhong, L.; Zolotukhin, S.; Srivastava, A.; Lewin, A.S.; et al. High-efficiency Transduction of the Mouse Retina by Tyrosine-mutant AAV Serotype Vectors. Mol. Ther. 2009, 17, 463–471. [Google Scholar] [CrossRef]
  171. Shen, X.; Xu, Y.; Bai, Z.; Ma, D.; Niu, Q.; Meng, J.; Fan, S.; Zhang, L.; Hao, Z.; Zhang, X.J. Transparenchymal renal pelvis injection of recombinant adeno-associated virus serotype 9 vectors is a practical approach for gene delivery in the kidney. Hum. Gene Ther. Methods 2018, 29, 251–258. [Google Scholar] [CrossRef]
  172. Konkalmatt, P.; Asico, L.; Feranil, J.; Jose, P.; Armando, I. Efficient in vivo gene transfer to murine renal cells using AAV9 vectors (912.7). FASEB J. 2014, 28, 917. [Google Scholar] [CrossRef]
  173. Schievenbusch, S.; Strack, I.; Scheffler, M.; Nischt, R.; Coutelle, O.; Hösel, M.; Hallek, M.; Fries, J.W.U.; Dienes, H.-P.; Odenthal, M.; et al. Combined Paracrine and Endocrine AAV9 mediated Expression of Hepatocyte Growth Factor for the Treatment of Renal Fibrosis. Mol. Ther. 2010, 18, 1302–1309. [Google Scholar] [CrossRef]
  174. Darbey, A.; Rebourcet, D.; Curley, M.; Kilcoyne, K.; Jeffery, N.; Reed, N.; Milne, L.; Roesl, C.; Brown, P.; Smith, L.B. A comparison of in vivo viral targeting systems identifies adeno-associated virus serotype 9 (AAV9) as an effective vector for genetic manipulation of Leydig cells in adult mice. Andrology 2021, 9, 460–473. [Google Scholar] [CrossRef]
  175. Limberis, M.P.; Wilson, J.M. Adeno-associated virus serotype 9 vectors transduce murine alveolar and nasal epithelia and can be readministered. Proc. Natl. Acad. Sci. USA 2006, 103, 12993–12998. [Google Scholar] [CrossRef] [Green Version]
  176. Bell, C.L.; Vandenberghe, L.H.; Bell, P.; Limberis, M.P.; Gao, G.-P.; Van Vliet, K.; Agbandje-McKenna, M.; Wilson, J.M. The AAV9 receptor and its modification to improve in vivo lung gene transfer in mice. J. Clin. Investig. 2011, 121, 2427–2435. [Google Scholar] [CrossRef] [Green Version]
  177. Vance, M.A.; Mitchell, A.; Samulski, R.J. AAV Biology, Infectivity and Therapeutic Use from Bench to Clinic. In Gene Therapy—Principles and Challenges; IntechOpen: London, UK, 2015. [Google Scholar] [CrossRef] [Green Version]
  178. Vandenberghe, L.H.; Miller, J.R.; Gao, G.-P.; Calcedo, R.; Wilson, J.M. 116. The Proposed AAV Serotypes 10 and 11 Serologically Cross-React with AAV8 and AAV4. Mol. Ther. 2006, 13, S47. [Google Scholar] [CrossRef]
  179. Klein, R.L.; Dayton, R.D.; Tatom, J.B.; Henderson, K.M.; Henning, P.P. AAV8, 9, Rh10, Rh43 Vector Gene Transfer in the Rat Brain: Effects of Serotype, Promoter and Purification Method. Mol. Ther. 2008, 16, 89–96. [Google Scholar] [CrossRef]
  180. Mori, S.; Takeuchi, T.; Enomoto, Y.; Kondo, K.; Sato, K.; Ono, F.; Sata, T.; Kanda, T. Tissue distribution of cynomolgus adeno-associated viruses AAV10, AAV11, and AAVcy.7 in naturally infected monkeys. Arch. Virol. 2008, 153, 375–380. [Google Scholar] [CrossRef]
  181. Polyak, S.; Mach, A.; Porvasnik, S.; Dixon, L.; Conlon, T.; Erger, K.E.; Acosta, A.; Wright, A.J.; Campbell-Thompson, M.; Zolotukhin, I.; et al. Identification of adeno-associated viral vectors suitable for intestinal gene delivery and modulation of experimental colitis. Am. J. Physiol.-Gastrointest. Liver Physiol. 2012, 302, G296–G308. [Google Scholar] [CrossRef] [Green Version]
  182. Singh, M.K.; Uehara, H.; Zhang, X.; Vashist, D.; Carroll, L.; Archer, B.; Huang, Y.; Ambati, B. Systemic Distribution of AAV10.COMPAng1 in the Ins2Akita mouse. Investig. Ophthalmol. Vis. Sci. 2016, 57, 5056. [Google Scholar]
  183. Tian, M.; Carroll, L.S.; Tang, L.; Uehara, H.; Westenfelder, C.; Ambati, B.K.; Huang, Y. Systemic AAV10. COMP-Ang1 rescues renal glomeruli and pancreatic islets in type 2 diabetic mice. BMJ Open Diabetes Res. Care 2020, 8, e000882. [Google Scholar] [CrossRef] [PubMed]
  184. Han, Z.; Luo, N.; Kou, J.; Li, L.; Ma, W.; Peng, S.; Xu, Z.; Zhang, W.; Qiu, Y.; Wu, Y. AAV11 permits efficient retrograde targeting of projection neurons. bioRxiv 2022. [Google Scholar] [CrossRef]
  185. Schmidt, M.; Voutetakis, A.; Afione, S.; Zheng, C.; Mandikian, D.; Chiorini, J.A. Adeno-Associated Virus Type 12 (AAV12): A Novel AAV Serotype with Sialic Acid- and Heparan Sulfate Proteoglycan-Independent Transduction Activity. J. Virol. 2008, 82, 1399–1406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Schmidt, M.; Voutetakis, A.; Afione, S.; Zheng, C.; Chiorini, J.A. 745. AAV12, Isolated from Vervet Monkey, Has Unique Tropism and Biological as Well as Neutralization Properties. Mol. Ther. 2006, 13, S288. [Google Scholar] [CrossRef]
  187. Quinn, K.; Quirion, M.R.; Lo, C.-Y.; Misplon, J.A.; Epstein, S.L.; Chiorini, J.A. Intranasal Administration of Adeno-associated Virus Type 12 (AAV12) Leads to Transduction of the Nasal Epithelia and Can Initiate Transgene-specific Immune Response. Mol. Ther. 2011, 19, 1990–1998. [Google Scholar] [CrossRef]
  188. Mietzsch, M.; Jose, A.; Chipman, P.; Bhattacharya, N.; Daneshparvar, N.; McKenna, R.; Agbandje-McKenna, M. Completion of the AAV Structural Atlas: Serotype Capsid Structures Reveals Clade-Specific Features. Viruses 2021, 13, 101. [Google Scholar] [CrossRef]
  189. Schmidt, M.; Govindasamy, L.; Afione, S.; Kaludov, N.; Agbandje-McKenna, M.; Chiorini, J.A. Molecular characterization of the heparin-dependent transduction domain on the capsid of a novel adeno-associated virus isolate, AAV(VR-942). J. Virol. 2008, 82, 8911–8916. [Google Scholar] [CrossRef] [Green Version]
  190. Huang, L.-Y.; Halder, S.; Agbandje-McKenna, M. Parvovirus glycan interactions. Curr. Opin. Virol. 2014, 7, 108–118. [Google Scholar] [CrossRef] [Green Version]
  191. Srivastava, A.J.C.; Insights, G.T. Advances and challenges in the use of recombinant AAV vectors for human gene therapy. Cell Gene Ther. Insights 2016, 2, 553–575. [Google Scholar] [CrossRef]
  192. Muzyczka, N.; Warrington, K.H., Jr. Custom adeno-associated virus capsids: The next generation of recombinant vectors with novel tropism. Hum. Gene Ther. 2005, 16, 408–416. [Google Scholar] [CrossRef]
  193. Zhang, L.; Rossi, A.; Lange, L.; Meumann, N.; Koitzsch, U.; Christie, K.; Nesbit, M.A.; Moore, C.B.T.; Hacker, U.T.; Morgan, M.; et al. Capsid Engineering Overcomes Barriers Toward Adeno-Associated Virus Vector-Mediated Transduction of Endothelial Cells. Hum. Gene Ther. 2019, 30, 1284–1296. [Google Scholar] [CrossRef]
  194. Ward, P.; Walsh, C.E. Chimeric AAV Cap sequences alter gene transduction. Virology 2009, 386, 237–248. [Google Scholar] [CrossRef] [Green Version]
  195. Yang, L.; Jiang, J.; Drouin, L.M.; Agbandje-McKenna, M.; Chen, C.; Qiao, C.; Pu, D.; Hu, X.; Wang, D.Z.; Li, J.; et al. A myocardium tropic adeno-associated virus (AAV) evolved by DNA shuffling and in vivo selection. Proc. Natl. Acad. Sci. USA 2009, 106, 3946–3951. [Google Scholar] [CrossRef] [Green Version]
  196. White, A.F.; Mazur, M.; Sorscher, E.J.; Zinn, K.R.; Ponnazhagan, S. Genetic modification of adeno-associated viral vector type 2 capsid enhances gene transfer efficiency in polarized human airway epithelial cells. Hum. Gene Ther. 2008, 19, 1407–1414. [Google Scholar] [CrossRef]
  197. Zhang, C.; Freistaedter, A.; Schmelas, C.; Gunkel, M.; Dao Thi, V.L.; Grimm, D. An RNA Interference/Adeno-Associated Virus Vector-Based Combinatorial Gene Therapy Approach Against Hepatitis E Virus. Hepatol. Commun. 2022, 6, 878–888. [Google Scholar] [CrossRef]
  198. Van Vliet, K.M.; Blouin, V.; Brument, N.; Agbandje-McKenna, M.; Snyder, R.O. The role of the adeno-associated virus capsid in gene transfer. Methods Mol. Biol. 2008, 437, 51–91. [Google Scholar] [CrossRef]
  199. Perabo, L.; Endell, J.; King, S.; Lux, K.; Goldnau, D.; Hallek, M.; Büning, H. Combinatorial engineering of a gene therapy vector: Directed evolution of adeno-associated virus. J. Gene Med. 2006, 8, 155–162. [Google Scholar] [CrossRef]
  200. Maheshri, N.; Koerber, J.T.; Kaspar, B.K.; Schaffer, D.V. Directed evolution of adeno-associated virus yields enhanced gene delivery vectors. Nat. Biotechnol. 2006, 24, 198–204. [Google Scholar] [CrossRef]
  201. Xiao, X.; deVlaminck, W.; Monahan, J. Adeno-associated virus (AAV) vectors for gene transfer. Adv. Drug Deliv. Rev. 1993, 12, 201–215. [Google Scholar] [CrossRef]
  202. Han, Z.; Zhong, L.; Maina, N.; Hu, Z.; Li, X.; Chouthai, N.S.; Bischof, D.; Weigel-Van Aken, K.A.; Slayton, W.B.; Yoder, M.C. Stable integration of recombinant adeno-associated virus vector genomes after transduction of murine hematopoietic stem cells. Hum. Gene Ther. 2008, 19, 267–278. [Google Scholar] [CrossRef]
  203. Tratschin, J.-D.; Miller, I.; Smith, M.; Carter, B.J.M.; Biology, C. Adeno-associated virus vector for high-frequency integration, expression, and rescue of genes in mammalian cells. Mol. Cell. Biol. 1985, 5, 3251–3260. [Google Scholar] [PubMed] [Green Version]
  204. Samulski, R.; Zhu, X.; Xiao, X.; Brook, J.; Housman, D.; Epstein, N.A.; Hunter, L. Targeted integration of adeno-associated virus (AAV) into human chromosome 19. EMBO J. 1991, 10, 3941–3950. [Google Scholar] [CrossRef]
  205. McCarty, D.M.; Young, S.M., Jr.; Samulski, R.J. Integration of adeno-associated virus (AAV) and recombinant AAV vectors. Annu. Rev. Genet. 2004, 38, 819–845. [Google Scholar] [CrossRef] [PubMed]
  206. Walsh, C.E.; Liu, J.M.; Xiao, X.; Young, N.S.; Nienhuis, A.W.; Samulski, R.J. Regulated high level expression of a human gamma-globin gene introduced into erythroid cells by an adeno-associated virus vector. Proc. Natl. Acad. Sci. USA 1992, 89, 7257–7261. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Winocour, E.; Callaham, M.; Huberman, E. Perturbation of the cell cycle by adeno-associated virus. Virology 1988, 167, 393–399. [Google Scholar] [CrossRef]
  208. Bantel-Schaal, U.; Stöhr, M.J. Influence of adeno-associated virus on adherence and growth properties of normal cells. J. Virol. 1992, 66, 773–779. [Google Scholar] [CrossRef] [Green Version]
  209. Samulski, R.J. Adeno-associated virus: Integration at a specific chromosomal locus. Curr. Opin. Genet. Dev. 1993, 3, 74–80. [Google Scholar] [CrossRef]
  210. Bijlani, S.; Pang, K.M.; Sivanandam, V.; Singh, A.; Chatterjee, S. The Role of Recombinant AAV in Precise Genome Editing. Front. Genome Ed. 2021, 3, 799722. [Google Scholar] [CrossRef]
  211. Wang, D.; Tai, P.W.; Gao, G. Adeno-associated virus vector as a platform for gene therapy delivery. Nat. Rev. Drug Discov. 2019, 18, 358–378. [Google Scholar] [CrossRef]
  212. Calcedo, R.; Wilson, J.M. Humoral immune response to AAV. Front. Immunol. 2013, 4, 341. [Google Scholar] [CrossRef] [Green Version]
  213. Hirsch, M.L.; Samulski, R.J. AAV-mediated gene editing via double-strand break repair. In Gene Correction; Springer: Berlin/Heidelberg, Germany, 2014; pp. 291–307. [Google Scholar]
  214. Gray, S.J.; Nagabhushan Kalburgi, S.; McCown, T.J.; Jude Samulski, R. Global CNS gene delivery and evasion of anti-AAV-neutralizing antibodies by intrathecal AAV administration in non-human primates. Gene Ther. 2013, 20, 450–459. [Google Scholar] [CrossRef] [Green Version]
  215. Ge, L.-J.; Yang, F.-H.; Li, W.; Wang, T.; Lin, Y.; Feng, J.; Chen, N.-H.; Jiang, M.; Wang, J.-H.; Hu, X.-T. In vivo neuroregeneration to treat ischemic stroke through NeuroD1 AAV-based gene therapy in adult non-human primates. Front. Cell Dev. Biol. 2020, 8, 590008. [Google Scholar] [CrossRef]
  216. Hu, C.; Lipshutz, G.S. AAV-based neonatal gene therapy for hemophilia A: Long-term correction and avoidance of immune responses in mice. Gene Ther. 2012, 19, 1166–1176. [Google Scholar] [CrossRef]
  217. Suzuki, J.; Hashimoto, K.; Xiao, R.; Vandenberghe, L.H.; Liberman, M.C. Cochlear gene therapy with ancestral AAV in adult mice: Complete transduction of inner hair cells without cochlear dysfunction. Sci. Rep. 2017, 7, 45524. [Google Scholar] [CrossRef] [Green Version]
  218. Akil, O.; Dyka, F.; Calvet, C.; Emptoz, A.; Lahlou, G.; Nouaille, S.; Boutet de Monvel, J.; Hardelin, J.-P.; Hauswirth, W.W.; Avan, P. Dual AAV-mediated gene therapy restores hearing in a DFNB9 mouse model. Proc. Natl. Acad. Sci. USA 2019, 116, 4496–4501. [Google Scholar] [CrossRef] [Green Version]
  219. Acland, G.M.; Aguirre, G.D.; Ray, J.; Zhang, Q.; Aleman, T.S.; Cideciyan, A.V.; Pearce-Kelling, S.E.; Anand, V.; Zeng, Y.; Maguire, A.M. Gene therapy restores vision in a canine model of childhood blindness. Nat. Genet. 2001, 28, 92–95. [Google Scholar] [CrossRef]
  220. Kodippili, K.; Hakim, C.H.; Pan, X.; Yang, H.T.; Yue, Y.; Zhang, Y.; Shin, J.-H.; Yang, N.N.; Duan, D. Dual AAV gene therapy for Duchenne muscular dystrophy with a 7-kb mini-dystrophin gene in the canine model. Hum. Gene Ther. 2018, 29, 299–311. [Google Scholar] [CrossRef]
  221. Gurda, B.L.; De Guilhem De Lataillade, A.; Bell, P.; Zhu, Y.; Yu, H.; Wang, P.; Bagel, J.; Vite, C.H.; Sikora, T.; Hinderer, C.; et al. Evaluation of AAV-mediated Gene Therapy for Central Nervous System Disease in Canine Mucopolysaccharidosis VII. Mol. Ther. J. Am. Soc. Gene Ther. 2016, 24, 206–216. [Google Scholar] [CrossRef] [Green Version]
  222. Gray-Edwards, H.L.; Maguire, A.S.; Salibi, N.; Ellis, L.E.; Voss, T.L.; Diffie, E.B.; Koehler, J.; Randle, A.N.; Taylor, A.R.; Brunson, B.L.; et al. 7T MRI predicts amelioration of neurodegeneration in the brain after AAV gene therapy. Mol. Ther.-Methods Clin. Dev. 2020, 17, 258–270. [Google Scholar] [CrossRef] [Green Version]
  223. Choudhury, S.R.; Fitzpatrick, Z.; Harris, A.F.; Maitland, S.A.; Ferreira, J.S.; Zhang, Y.; Ma, S.; Sharma, R.B.; Gray-Edwards, H.L.; Johnson, J.A. In vivo selection yields AAV-B1 capsid for central nervous system and muscle gene therapy. Mol. Ther. 2016, 24, 1247–1257. [Google Scholar] [CrossRef] [Green Version]
  224. McCown, T.J. Adeno-associated virus (AAV) vectors in the CNS. Curr. Gene Ther. 2005, 5, 333–338. [Google Scholar] [CrossRef] [PubMed]
  225. Kaplitt, M.G.; Leone, P.; Samulski, R.J.; Xiao, X.; Pfaff, D.W.; O’Malley, K.L.; During, M.J. Long-term gene expression and phenotypic correction using adeno-associated virus vectors in the mammalian brain. Nat. Genet. 1994, 8, 148–154. [Google Scholar] [CrossRef] [PubMed]
  226. McElroy, A.; Sena-Esteves, M.; Arjomandnejad, M.; Keeler, A.M.; Gray-Edwards, H.L. Redosing Adeno-Associated Virus Gene Therapy to the Central Nervous System. Hum. Gene Ther. 2022, 33, 889–892. [Google Scholar] [CrossRef] [PubMed]
  227. Bartus, R.T.; Weinberg, M.S.; Samulski, R.J. Parkinson’s Disease Gene Therapy: Success by Design Meets Failure by Efficacy. Mol. Ther. 2014, 22, 487–497. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Marks, W.J.; Bartus, R.T.; Siffert, J.; Davis, C.S.; Lozano, A.; Boulis, N.; Vitek, J.; Stacy, M.; Turner, D.; Verhagen, L.; et al. Gene delivery of AAV2-neurturin for Parkinson’s disease: A double-blind, randomised, controlled trial. Lancet Neurol. 2010, 9, 1164–1172. [Google Scholar] [CrossRef]
  229. Kaplitt, M.G.; Feigin, A.; Tang, C.; Fitzsimons, H.L.; Mattis, P.; Lawlor, P.A.; Bland, R.J.; Young, D.; Strybing, K.; Eidelberg, D.; et al. Safety and tolerability of gene therapy with an adeno-associated virus (AAV) borne GAD gene for Parkinson’s disease: An open label, phase I trial. Lancet 2007, 369, 2097–2105. [Google Scholar] [CrossRef]
  230. LeWitt, P.A.; Rezai, A.R.; Leehey, M.A.; Ojemann, S.G.; Flaherty, A.W.; Eskandar, E.N.; Kostyk, S.K.; Thomas, K.; Sarkar, A.; Siddiqui, M.S.; et al. AAV2-GAD gene therapy for advanced Parkinson’s disease: A double-blind, sham-surgery controlled, randomised trial. Lancet Neurol. 2011, 10, 309–319. [Google Scholar] [CrossRef]
  231. Bartus, R.T.; Baumann, T.L.; Siffert, J.; Herzog, C.D.; Alterman, R.; Boulis, N.; Turner, D.A.; Stacy, M.; Lang, A.E.; Lozano, A.M.; et al. Safety/feasibility of targeting the substantia nigra with AAV2-neurturin in Parkinson patients. Neurology 2013, 80, 1698–1701. [Google Scholar] [CrossRef] [Green Version]
  232. Christine, C.W.; Starr, P.A.; Larson, P.S.; Eberling, J.L.; Jagust, W.J.; Hawkins, R.A.; VanBrocklin, H.F.; Wright, J.F.; Bankiewicz, K.S.; Aminoff, M.J. Safety and tolerability of putaminal AADC gene therapy for Parkinson disease. Neurology 2009, 73, 1662–1669. [Google Scholar] [CrossRef] [Green Version]
  233. Mittermeyer, G.; Christine, C.W.; Rosenbluth, K.H.; Baker, S.L.; Starr, P.; Larson, P.; Kaplan, P.L.; Forsayeth, J.; Aminoff, M.J.; Bankiewicz, K.S. Long-term evaluation of a phase 1 study of AADC gene therapy for Parkinson’s disease. Hum. Gene Ther. 2012, 23, 377–381. [Google Scholar] [CrossRef] [Green Version]
  234. Chien, Y.H.; Lee, N.C.; Tseng, S.H.; Tai, C.H.; Muramatsu, S.I.; Byrne, B.J.; Hwu, W.L. Efficacy and safety of AAV2 gene therapy in children with aromatic L-amino acid decarboxylase deficiency: An open-label, phase 1/2 trial. Lancet Child Adolesc. Health 2017, 1, 265–273. [Google Scholar] [CrossRef] [Green Version]
  235. Kolb, S.J.; Kissel, J.T. Spinal muscular atrophy: A timely review. Arch. Neurol. 2011, 68, 979–984. [Google Scholar] [CrossRef] [Green Version]
  236. Aguti, S.; Malerba, A.; Zhou, H. The progress of AAV-mediated gene therapy in neuromuscular disorders. Expert Opin. Biol. Ther. 2018, 18, 681–693. [Google Scholar] [CrossRef]
  237. Spahiu, L.; Behluli, E.; Peterlin, B.; Nefic, H.; Hadziselimovic, R.; Liehr, T.; Temaj, G. Mucopolysaccharidosis III: Molecular basis and treatment. Pediatr. Endocrinol. Diabetes Metab. 2021, 27, 201–208. [Google Scholar] [CrossRef]
  238. Watson, G.L.; Sayles, J.N.; Chen, C.; Elliger, S.S.; Elliger, C.A.; Raju, N.R.; Kurtzman, G.J.; Podsakoff, G.M. Treatment of lysosomal storage disease in MPS VII mice using a recombinant adeno-associated virus. Gene Ther. 1998, 5, 1642–1649. [Google Scholar] [CrossRef] [Green Version]
  239. Skorupa, A.F.; Fisher, K.J.; Wilson, J.M.; Parente, M.K.; Wolfe, J.H. Sustained production of beta-glucuronidase from localized sites after AAV vector gene transfer results in widespread distribution of enzyme and reversal of lysosomal storage lesions in a large volume of brain in mucopolysaccharidosis VII mice. Exp. Neurol. 1999, 160, 17–27. [Google Scholar] [CrossRef]
  240. Daly, T.M.; Vogler, C.; Levy, B.; Haskins, M.E.; Sands, M.S. Neonatal gene transfer leads to widespread correction of pathology in a murine model of lysosomal storage disease. Proc. Natl. Acad. Sci. USA 1999, 96, 2296–2300. [Google Scholar] [CrossRef] [Green Version]
  241. Marcó, S.; Haurigot, V.; Jaén, M.L.; Ribera, A.; Sánchez, V.; Molas, M.; Garcia, M.; León, X.; Roca, C.; Sánchez, X.; et al. Seven-year follow-up of durability and safety of AAV CNS gene therapy for a lysosomal storage disorder in a large animal. Mol. Ther.-Methods Clin. Dev. 2021, 23, 370–389. [Google Scholar] [CrossRef]
  242. Shaimardanova, A.A.; Chulpanova, D.S.; Solovyeva, V.V.; Mullagulova, A.I.; Kitaeva, K.V.; Allegrucci, C.; Rizvanov, A.A. Metachromatic Leukodystrophy: Diagnosis, Modeling, and Treatment Approaches. Front. Med. 2020, 7, 576221. [Google Scholar] [CrossRef]
  243. Piguet, F.; Sondhi, D.; Piraud, M.; Fouquet, F.; Hackett, N.R.; Ahouansou, O.; Vanier, M.-T.; Bieche, I.; Aubourg, P.; Crystal, R.G. Correction of brain oligodendrocytes by AAVrh. 10 intracerebral gene therapy in metachromatic leukodystrophy mice. Hum. Gene Ther. 2012, 23, 903–914. [Google Scholar] [CrossRef] [Green Version]
  244. Kurai, T.; Hisayasu, S.; Kitagawa, R.; Migita, M.; Suzuki, H.; Hirai, Y.; Shimada, T. AAV1 mediated co-expression of formylglycine-generating enzyme and arylsulfatase a efficiently corrects sulfatide storage in a mouse model of metachromatic leukodystrophy. Mol. Ther. 2007, 15, 38–43. [Google Scholar] [CrossRef] [PubMed]
  245. Miyake, N.; Miyake, K.; Asakawa, N.; Yamamoto, M.; Shimada, T. Long-term correction of biochemical and neurological abnormalities in MLD mice model by neonatal systemic injection of an AAV serotype 9 vector. Gene Ther. 2014, 21, 427–433. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. Sevin, C.; Benraiss, A.; Van Dam, D.; Bonnin, D.; Nagels, G.; Verot, L.; Laurendeau, I.; Vidaud, M.; Gieselmann, V.; Vanier, M. Intracerebral adeno-associated virus-mediated gene transfer in rapidly progressive forms of metachromatic leukodystrophy. Hum. Mol. Genet. 2006, 15, 53–64. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  247. Shaimardanova, A.A.; Chulpanova, D.S.; Solovyeva, V.V.; Aimaletdinov, A.M.; Rizvanov, A.A. Functionality of a bicistronic construction containing HEXA and HEXB genes encoding β-hexosaminidase A for cell-mediated therapy of GM2 gangliosidoses. Neural Regen. Res. 2022, 17, 122–129. [Google Scholar] [CrossRef]
  248. Taghian, T.; Marosfoi, M.G.; Puri, A.S.; Cataltepe, O.I.; King, R.M.; Diffie, E.B.; Maguire, A.S.; Martin, D.R.; Fernau, D.; Batista, A.R.; et al. A Safe and Reliable Technique for CNS Delivery of AAV Vectors in the Cisterna Magna. Mol. Ther. J. Am. Soc. Gene Ther. 2020, 28, 411–421. [Google Scholar] [CrossRef]
  249. Solovyeva, V.V.; Shaimardanova, A.A.; Chulpanova, D.S.; Kitaeva, K.V.; Chakrabarti, L.; Rizvanov, A.A. New Approaches to Tay-Sachs Disease Therapy. Front. Physiol. 2018, 9, 1663. [Google Scholar] [CrossRef] [Green Version]
  250. Cachón-González, M.B.; Wang, S.Z.; Lynch, A.; Ziegler, R.; Cheng, S.H.; Cox, T.M. Effective gene therapy in an authentic model of Tay-Sachs-related diseases. Proc. Natl. Acad. Sci. USA 2006, 103, 10373–10378. [Google Scholar] [CrossRef] [Green Version]
  251. Gray-Edwards, H.L.; Randle, A.N.; Maitland, S.A.; Benatti, H.R.; Hubbard, S.M.; Canning, P.F.; Vogel, M.B.; Brunson, B.L.; Hwang, M.; Ellis, L.E.; et al. Adeno-Associated Virus Gene Therapy in a Sheep Model of Tay-Sachs Disease. Hum. Gene Ther. 2018, 29, 312–326. [Google Scholar] [CrossRef]
  252. Rockwell, H.E.; McCurdy, V.J.; Eaton, S.C.; Wilson, D.U.; Johnson, A.K.; Randle, A.N.; Bradbury, A.M.; Gray-Edwards, H.L.; Baker, H.J.; Hudson, J.A.; et al. AAV-mediated gene delivery in a feline model of Sandhoff disease corrects lysosomal storage in the central nervous system. ASN Neuro 2015, 7, 1759091415569908. [Google Scholar] [CrossRef]
  253. Sevin, C.; Roujeau, T.; Cartier, N.; Baugnon, T.; Adamsbaum, C.; Piraud, M.; Martino, S.; Mouiller, P.; Couzinié, C.; Bellesme, C. Intracerebral Gene Therapy in Children with Metachromatic Leukodystrophy: Results of a Phase I/II Trial. Mol. Genet. Metab. 2018, 123, S129. [Google Scholar] [CrossRef]
  254. Tenenbaum, L.; Chtarto, A.; Lehtonen, E.; Velu, T.; Brotchi, J.; Levivier, M. Recombinant AAV-mediated gene delivery to the central nervous system. J. Gene Med. 2004, 6, S212–S222. [Google Scholar] [CrossRef]
  255. Surendran, S.; Michals-Matalon, K.; Quast, M.J.; Tyring, S.K.; Wei, J.; Ezell, E.L.; Matalon, R. Canavan disease: A monogenic trait with complex genomic interaction. Mol. Genet. Metab. 2003, 80, 74–80. [Google Scholar] [CrossRef]
  256. Matalon, R.; Surendran, S.; Rady, P.L.; Quast, M.J.; Campbell, G.A.; Matalon, K.M.; Tyring, S.K.; Wei, J.; Peden, C.S.; Ezell, E.L.; et al. Adeno-associated virus-mediated aspartoacylase gene transfer to the brain of knockout mouse for canavan disease. Mol. Ther. J. Am. Soc. Gene Ther. 2003, 7, 580–587. [Google Scholar] [CrossRef]
  257. Janson, C.; McPhee, S.; Bilaniuk, L.; Haselgrove, J.; Testaiuti, M.; Freese, A.; Wang, D.J.; Shera, D.; Hurh, P.; Rupin, J.; et al. Clinical protocol. Gene therapy of Canavan disease: AAV-2 vector for neurosurgical delivery of aspartoacylase gene (ASPA) to the human brain. Hum. Gene Ther. 2002, 13, 1391–1412. [Google Scholar] [CrossRef] [Green Version]
  258. rAAV-Olig001-ASPA Gene Therapy for Treatment of Children With Typical Canavan Disease. Available online: https://ClinicalTrials.gov/show/NCT04833907 (accessed on 1 December 2022).
  259. Canavan-Single Patient IND. Available online: https://ClinicalTrials.gov/show/NCT05317780 (accessed on 1 December 2022).
  260. A Study of AAV9 Gene Therapy in Participants With Canavan Disease. Available online: https://ClinicalTrials.gov/show/NCT04998396 (accessed on 1 December 2022).
  261. Nasir, G.; Chopra, R.; Elwood, F.; Ahmed, S.S. Krabbe Disease: Prospects of Finding a Cure Using AAV Gene Therapy. Front. Med. 2021, 8, 760236. [Google Scholar] [CrossRef]
  262. Rafi, M.A.; Rao, H.Z.; Passini, M.A.; Curtis, M.; Vanier, M.T.; Zaka, M.; Luzi, P.; Wolfe, J.H.; Wenger, D.A. AAV-mediated expression of galactocerebrosidase in brain results in attenuated symptoms and extended life span in murine models of globoid cell leukodystrophy. Mol. Ther. 2005, 11, 734–744. [Google Scholar] [CrossRef]
  263. Lin, D.-S.; Hsiao, C.-D.; Liau, I.; Lin, S.-P.; Chiang, M.-F.; Chuang, C.-K.; Wang, T.-J.; Wu, T.-Y.; Jian, Y.-R.; Huang, S.-F.; et al. CNS-targeted AAV5 gene transfer results in global dispersal of vector and prevention of morphological and function deterioration in CNS of globoid cell leukodystrophy mouse model. Mol. Genet. Metab. 2011, 103, 367–377. [Google Scholar] [CrossRef]
  264. Lin, D.; Fantz, C.R.; Levy, B.; Rafi, M.A.; Vogler, C.; Wenger, D.A.; Sands, M.S. AAV2/5 vector expressing galactocerebrosidase ameliorates CNS disease in the murine model of globoid-cell leukodystrophy more efficiently than AAV2. Mol. Ther. 2005, 12, 422–430. [Google Scholar] [CrossRef]
  265. Karumuthil-Melethil, S.; Marshall, M.S.; Heindel, C.; Jakubauskas, B.; Bongarzone, E.R.; Gray, S.J. 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. 2016, 94, 1138–1151. [Google Scholar] [CrossRef] [Green Version]
  266. Bradbury, A.M.; Rafi, M.A.; Bagel, J.H.; Brisson, B.K.; Marshall, M.S.; Pesayco Salvador, J.; Jiang, X.; Swain, G.P.; Prociuk, M.L.; ODonnell, P.A. AAVrh10 gene therapy ameliorates central and peripheral nervous system disease in canine globoid cell leukodystrophy (Krabbe disease). Hum. Gene Ther. 2018, 29, 785–801. [Google Scholar] [CrossRef]
  267. Bradbury, A.M.; Bagel, J.H.; Nguyen, D.; Lykken, E.A.; Salvador, J.P.; Jiang, X.; Swain, G.P.; Assenmacher, C.A.; Hendricks, I.J.; Miyadera, K. Krabbe disease successfully treated via monotherapy of intrathecal gene therapy. J. Clin. Investig. 2020, 130, 4906–4920. [Google Scholar] [CrossRef] [PubMed]
  268. Hori, T.; Fukutome, M.; Maejima, C.; Matsushima, H.; Kobayashi, K.; Kitazawa, S.; Kitahara, R.; Kitano, K.; Kobayashi, K.; Moritoh, S.; et al. Gene delivery to cone photoreceptors by subretinal injection of rAAV2/6 in the mouse retina. Biochem. Biophys. Res. Commun. 2019, 515, 222–227. [Google Scholar] [CrossRef] [PubMed]
  269. Le Meur, G.; Lebranchu, P.; Billaud, F.; Adjali, O.; Schmitt, S.; Bézieau, S.; Péréon, Y.; Valabregue, R.; Ivan, C.; Darmon, C.; et al. Safety and Long-Term Efficacy of AAV4 Gene Therapy in Patients with RPE65 Leber Congenital Amaurosis. Mol. Ther. J. Am. Soc. Gene Ther. 2018, 26, 256–268. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  270. Le Meur, G.; Lebranchu, P.; Pereon, Y.; Billaud, F.; Ivan, C.; Chauveau, C.; Rolling, F.; Weber, M. Efficacy and safety of gene therapy with AAV4 in childhood blindness due to rpe65 mutations. Acta Ophthalmol. 2012, 90. [Google Scholar] [CrossRef]
  271. Trapani, I.; Auricchio, A. Seeing the Light after 25 Years of Retinal Gene Therapy. Trends Mol. Med. 2018, 24, 669–681. [Google Scholar] [CrossRef]
  272. MacLaren, R.E.; Groppe, M.; Barnard, A.R.; Cottriall, C.L.; Tolmachova, T.; Seymour, L.; Clark, K.R.; During, M.J.; Cremers, F.P.M.; Black, G.C.M.; et al. Retinal gene therapy in patients with choroideremia: Initial findings from a phase 1/2 clinical trial. Lancet 2014, 383, 1129–1137. [Google Scholar] [CrossRef] [Green Version]
  273. Faynus, M.A.; Clegg, D.O. Chapter 7—Modeling inherited retinal dystrophies using induced pluripotent stem cells. In Current Progress in iPSC Disease Modeling; Birbrair, A., Ed.; Academic Press: Cambridge, MS, USA, 2022; pp. 157–184. [Google Scholar] [CrossRef]
  274. Russell, S.; Bennett, J.; Wellman, J.A.; Chung, D.C.; Yu, Z.-F.; Tillman, A.; Wittes, J.; Pappas, J.; Elci, O.; McCague, S.; et al. Efficacy and safety of voretigene neparvovec (AAV2-hRPE65v2) in patients with RPE65-mediated inherited retinal dystrophy: A randomised, controlled, open-label, phase 3 trial. Lancet 2017, 390, 849–860. [Google Scholar] [CrossRef]
  275. Bouquet, C.; Vignal Clermont, C.; Galy, A.; Fitoussi, S.; Blouin, L.; Munk, M.R.; Valero, S.; Meunier, S.; Katz, B.; Sahel, J.A.; et al. Immune Response and Intraocular Inflammation in Patients With Leber Hereditary Optic Neuropathy Treated With Intravitreal Injection of Recombinant Adeno-Associated Virus 2 Carrying the ND4 Gene: A Secondary Analysis of a Phase 1/2 Clinical Trial. JAMA Ophthalmol. 2019, 137, 399–406. [Google Scholar] [CrossRef] [Green Version]
  276. Feuer, W.J.; Schiffman, J.C.; Davis, J.L.; Porciatti, V.; Gonzalez, P.; Koilkonda, R.D.; Yuan, H.; Lalwani, A.; Lam, B.L.; Guy, J. Gene Therapy for Leber Hereditary Optic Neuropathy: Initial Results. Ophthalmology 2016, 123, 558–570. [Google Scholar] [CrossRef] [Green Version]
  277. Cehajic-Kapetanovic, J.; Xue, K.; Martinez-Fernandez de la Camara, C.; Nanda, A.; Davies, A.; Wood, L.J.; Salvetti, A.P.; Fischer, M.D.; Aylward, J.W.; Barnard, A.R.; et al. Initial results from a first-in-human gene therapy trial on X-linked retinitis pigmentosa caused by mutations in RPGR. Nat. Med. 2020, 26, 354–359. [Google Scholar] [CrossRef]
  278. Cukras, C.; Wiley, H.E.; Jeffrey, B.G.; Sen, H.N.; Turriff, A.; Zeng, Y.; Vijayasarathy, C.; Marangoni, D.; Ziccardi, L.; Kjellstrom, S.; et al. Retinal AAV8-RS1 Gene Therapy for X-Linked Retinoschisis: Initial Findings from a Phase I/IIa Trial by Intravitreal Delivery. Mol. Ther. 2018, 26, 2282–2294. [Google Scholar] [CrossRef] [Green Version]
  279. Mishra, A.; Vijayasarathy, C.; Cukras, C.A.; Wiley, H.E.; Sen, H.N.; Zeng, Y.; Wei, L.L.; Sieving, P.A. Immune function in X-linked retinoschisis subjects in an AAV8-RS1 phase I/IIa gene therapy trial. Mol. Ther. 2021, 29, 2030–2040. [Google Scholar] [CrossRef]
  280. Fischer, M.D.; Michalakis, S.; Wilhelm, B.; Zobor, D.; Muehlfriedel, R.; Kohl, S.; Weisschuh, N.; Ochakovski, G.A.; Klein, R.; Schoen, C.; et al. Safety and Vision Outcomes of Subretinal Gene Therapy Targeting Cone Photoreceptors in Achromatopsia: A Nonrandomized Controlled Trial. JAMA Ophthalmol. 2020, 138, 643–651. [Google Scholar] [CrossRef]
  281. Kahle, N.A.; Peters, T.; Zobor, D.; Kuehlewein, L.; Kohl, S.; Zhour, A.; Werner, A.; Seitz, I.P.; Sothilingam, V.; Michalakis, S.; et al. Development of Methodology and Study Protocol: Safety and Efficacy of a Single Subretinal Injection of rAAV.hCNGA3 in Patients with CNGA3-Linked Achromatopsia Investigated in an Exploratory Dose-Escalation Trial. Hum. Gene Ther. Clin. Dev. 2018, 29, 121–131. [Google Scholar] [CrossRef]
  282. Lalwani, A.K.; Walsh, B.J.; Reilly, P.G.; Muzyczka, N.; Mhatre, A.N. Development of in vivo gene therapy for hearing disorders: Introduction of adeno-associated virus into the cochlea of the guinea pig. Gene Ther. 1996, 3, 588–592. [Google Scholar]
  283. Sondhi, D.; Stiles, K.M.; De, B.P.; Crystal, R.G. Genetic Modification of the Lung Directed Toward Treatment of Human Disease. Hum. Gene Ther. 2017, 28, 3–84. [Google Scholar] [CrossRef]
  284. McLachlan, G.; Alton, E.; Boyd, A.C.; Clarke, N.K.; Davies, J.C.; Gill, D.R.; Griesenbach, U.; Hickmott, J.W.; Hyde, S.C.; Miah, K.M.; et al. Progress in Respiratory Gene Therapy. Hum. Gene Ther. 2022, 33, 893–912. [Google Scholar] [CrossRef]
  285. Burney, T.J.; Davies, J.C. Gene therapy for the treatment of cystic fibrosis. Appl. Clin. Genet. 2012, 5, 29. [Google Scholar]
  286. Flotte, T.R.; Afione, S.A.; Conrad, C.; McGrath, S.A.; Solow, R.; Oka, H.; Zeitlin, P.L.; Guggino, W.B.; Carter, B.J. Stable in vivo expression of the cystic fibrosis transmembrane conductance regulator with an adeno-associated virus vector. Proc. Natl. Acad. Sci. USA 1993, 90, 10613–10617. [Google Scholar] [CrossRef] [Green Version]
  287. Guggino, W.B.; Cebotaru, L. Gene therapy for cystic fibrosis paved the way for the use of adeno-associated virus in gene therapy. Hum. Gene Ther. 2020, 31, 538–541. [Google Scholar] [CrossRef]
  288. Wagner, J.A.; Reynolds, T.; Moran, M.L.; Moss, R.B.; Wine, J.J.; Flotte, T.R.; Gardner, P. Efficient and persistent gene transfer of AAV-CFTR in maxillary sinus. Lancet 1998, 351, 1702–1703. [Google Scholar] [CrossRef] [PubMed]
  289. Wagner, J.A.; Nepomuceno, I.B.; Messner, A.H.; Moran, M.L.; Batson, E.P.; Dimiceli, S.; Brown, B.W.; Desch, J.K.; Norbash, A.M.; Conrad, C.K. A phase II, double-blind, randomized, placebo-controlled clinical trial of tgAAVCF using maxillary sinus delivery in patients with cystic fibrosis with antrostomies. Hum. Gene Ther. 2002, 13, 1349–1359. [Google Scholar] [CrossRef] [PubMed]
  290. Aitken, M.; Moss, R.; Waltz, D.; Dovey, M.; Tonelli, M.; McNamara, S.; Gibson, R.; Ramsey, B.; Carter, B.; Reynolds, T. A phase I study of aerosolized administration of tgAAVCF to cystic fibrosis subjects with mild lung disease. Hum. Gene Ther. 2001, 12, 1907–1916. [Google Scholar] [CrossRef] [PubMed]
  291. Flotte, T.R.; Zeitlin, P.L.; Reynolds, T.C.; Heald, A.E.; Pedersen, P.; Beck, S.; Conrad, C.K.; Brass-Ernst, L.; Humphries, M.; Sullivan, K. Phase I trial of intranasal and endobronchial administration of a recombinant adeno-associated virus serotype 2 (rAAV2)-CFTR vector in adult cystic fibrosis patients: A two-part clinical study. Hum. Gene Ther. 2003, 14, 1079–1088. [Google Scholar] [CrossRef]
  292. Moss, R.B.; Rodman, D.; Spencer, L.T.; Aitken, M.L.; Zeitlin, P.L.; Waltz, D.; Milla, C.; Brody, A.S.; Clancy, J.P.; Ramsey, B. Repeated adeno-associated virus serotype 2 aerosol-mediated cystic fibrosis transmembrane regulator gene transfer to the lungs of patients with cystic fibrosis: A multicenter, double-blind, placebo-controlled trial. Chest 2004, 125, 509–521. [Google Scholar] [CrossRef] [Green Version]
  293. Moss, R.B.; Milla, C.; Colombo, J.; Accurso, F.; Zeitlin, P.L.; Clancy, J.P.; Spencer, L.T.; Pilewski, J.; Waltz, D.A.; Dorkin, H.L. Repeated aerosolized AAV-CFTR for treatment of cystic fibrosis: A randomized placebo-controlled phase 2B trial. Hum. Gene Ther. 2007, 18, 726–732. [Google Scholar] [CrossRef]
  294. Kurosaki, F.; Uchibori, R.; Mato, N.; Sehara, Y.; Saga, Y.; Urabe, M.; Mizukami, H.; Sugiyama, Y.; Kume, A. Optimization of adeno-associated virus vector-mediated gene transfer to the respiratory tract. Gene Ther. 2017, 24, 290–297. [Google Scholar] [CrossRef]
  295. Cooney, A.L.; McCray, P.B.; Sinn, P.L. Cystic Fibrosis Gene Therapy: Looking Back, Looking Forward. Genes 2018, 9, 538. [Google Scholar] [CrossRef] [Green Version]
  296. Stiles, K.M.; Sondhi, D.; Kaminsky, S.M.; De, B.P.; Rosenberg, J.B.; Crystal, R.G. Intrapleural Gene Therapy for Alpha-1 Antitrypsin Deficiency-Related Lung Disease. Chronic Obstr. Pulm. Dis. 2018, 5, 244–257. [Google Scholar] [CrossRef]
  297. Song, S.; Embury, J.; Laipis, P.; Berns, K.; Crawford, J.; Flotte, T. Stable therapeutic serum levels of human alpha-1 antitrypsin (AAT) after portal vein injection of recombinant adeno-associated virus (rAAV) vectors. Gene Ther. 2001, 8, 1299–1306. [Google Scholar] [CrossRef] [Green Version]
  298. Conlon, T.J.; Cossette, T.; Erger, K.; Choi, Y.-K.; Clarke, T.; Scott-Jorgensen, M.; Song, S.; Campbell-Thompson, M.; Crawford, J.; Flotte, T.R. Efficient Hepatic Delivery and Expression from a Recombinant Adeno-associated Virus 8 Pseudotyped α1-Antitrypsin Vector. Mol. Ther. 2005, 12, 867–875. [Google Scholar] [CrossRef]
  299. De, B.; Heguy, A.; Leopold, P.L.; Wasif, N.; Korst, R.J.; Hackett, N.R.; Crystal, R.G. Intrapleural administration of a serotype 5 adeno-associated virus coding for α1-antitrypsin mediates persistent, high lung and serum levels of α1-antitrypsin. Mol. Ther. 2004, 10, 1003–1010. [Google Scholar] [CrossRef]
  300. Chulay, J.D.; Ye, G.J.; Thomas, D.L.; Knop, D.R.; Benson, J.M.; Hutt, J.A.; Wang, G.; Humphries, M.; Flotte, T.R. Preclinical evaluation of a recombinant adeno-associated virus vector expressing human alpha-1 antitrypsin made using a recombinant herpes simplex virus production method. Hum. Gene Ther. 2011, 22, 155–165. [Google Scholar] [CrossRef]
  301. Limberis, M.P.; Vandenberghe, L.H.; Zhang, L.; Pickles, R.J.; Wilson, J.M. Transduction Efficiencies of Novel AAV Vectors in Mouse Airway Epithelium In Vivo and Human Ciliated Airway Epithelium In Vitro. Mol. Ther. 2009, 17, 294–301. [Google Scholar] [CrossRef]
  302. Liqun Wang, R.; McLaughlin, T.; Cossette, T.; Tang, Q.; Foust, K.; Campbell-Thompson, M.; Martino, A.; Cruz, P.; Loiler, S.; Mueller, C.; et al. Recombinant AAV Serotype and Capsid Mutant Comparison for Pulmonary Gene Transfer of α-1-Antitrypsin Using Invasive and Noninvasive Delivery. Mol. Ther. 2009, 17, 81–87. [Google Scholar] [CrossRef]
  303. Chiuchiolo, M.J.; Kaminsky, S.M.; Sondhi, D.; Hackett, N.R.; Rosenberg, J.B.; Frenk, E.Z.; Hwang, Y.; Van de Graaf, B.G.; Hutt, J.A.; Wang, G.; et al. Intrapleural administration of an AAVrh.10 vector coding for human α1-antitrypsin for the treatment of α1-antitrypsin deficiency. Hum. Gene Ther. Clin. Dev. 2013, 24, 161–173. [Google Scholar] [CrossRef] [Green Version]
  304. De, B.P.; Heguy, A.; Hackett, N.R.; Ferris, B.; Leopold, P.L.; Lee, J.; Pierre, L.; Gao, G.; Wilson, J.M.; Crystal, R.G. High Levels of Persistent Expression of α1-Antitrypsin Mediated by the Nonhuman Primate Serotype rh.10 Adeno-associated Virus Despite Preexisting Immunity to Common Human Adeno-associated Viruses. Mol. Ther. 2006, 13, 67–76. [Google Scholar] [CrossRef]
  305. Flotte, T.R.; Conlon, T.J.; Poirier, A.; Campbell-Thompson, M.; Byrne, B.J. Preclinical characterization of a recombinant adeno-associated virus type 1-pseudotyped vector demonstrates dose-dependent injection site inflammation and dissemination of vector genomes to distant sites. Hum. Gene Ther. 2007, 18, 245–256. [Google Scholar] [CrossRef]
  306. Zavorotinskaya, T.; Tomkinson, A.; Murphy, J.E. Treatment of experimental asthma by long-term gene therapy directed against IL-4 and IL-13. Mol. Ther. 2003, 7, 155–162. [Google Scholar] [CrossRef]
  307. Kang, M.H.; van Lieshout, L.P.; Xu, L.; Domm, J.M.; Vadivel, A.; Renesme, L.; Mühlfeld, C.; Hurskainen, M.; Mižíková, I.; Pei, Y. A lung tropic AAV vector improves survival in a mouse model of surfactant B deficiency. Nat. Commun. 2020, 11, 3929. [Google Scholar] [CrossRef]
  308. Cooney, A.L.; Wambach, J.A.; Sinn, P.L.; McCray, P.B., Jr. Gene therapy potential for genetic disorders of surfactant dysfunction. Front. Genome Ed. 2021, 3, 43. [Google Scholar] [CrossRef] [PubMed]
  309. Guse, K.; Suzuki, M.; Sule, G.; Bertin, T.K.; Tyynismaa, H.; Ahola-Erkkilä, S.; Palmer, D.; Suomalainen, A.; Ng, P.; Cerullo, V.; et al. Capsid-modified adenoviral vectors for improved muscle-directed gene therapy. Hum. Gene Ther. 2012, 23, 1065–1070. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  310. Tang, Y.; Cummins, J.; Huard, J.; Wang, B. AAV-directed muscular dystrophy gene therapy. Expert Opin. Biol. Ther. 2010, 10, 395–408. [Google Scholar] [CrossRef] [PubMed]
  311. Kota, J.; Handy, C.R.; Haidet, A.M.; Montgomery, C.L.; Eagle, A.; Rodino-Klapac, L.R.; Tucker, D.; Shilling, C.J.; Therlfall, W.R.; Walker, C.M.; et al. Follistatin gene delivery enhances muscle growth and strength in nonhuman primates. Sci. Transl. Med. 2009, 1, 6ra15. [Google Scholar] [CrossRef] [Green Version]
  312. Gregorevic, P.; Allen, J.M.; Minami, E.; Blankinship, M.J.; Haraguchi, M.; Meuse, L.; Finn, E.; Adams, M.E.; Froehner, S.C.; Murry, C.E.; et al. rAAV6-microdystrophin preserves muscle function and extends lifespan in severely dystrophic mice. Nat. Med. 2006, 12, 787–789. [Google Scholar] [CrossRef]
  313. Wang, B.; Li, J.; Fu, F.H.; Xiao, X. Systemic human minidystrophin gene transfer improves functions and life span of dystrophin and dystrophin/utrophin-deficient mice. J. Orthop. Res. Off. Publ. Orthop. Res. Soc. 2009, 27, 421–426. [Google Scholar] [CrossRef]
  314. Zhu, T.; Zhou, L.; Mori, S.; Wang, Z.; McTiernan, C.F.; Qiao, C.; Chen, C.; Wang, D.W.; Li, J.; Xiao, X. Sustained Whole-Body Functional Rescue in Congestive Heart Failure and Muscular Dystrophy Hamsters by Systemic Gene Transfer. Circulation 2005, 112, 2650–2659. [Google Scholar] [CrossRef] [Green Version]
  315. Pacak, C.A.; Walter, G.A.; Gaidosh, G.; Bryant, N.; Lewis, M.A.; Germain, S.; Mah, C.S.; Campbell, K.P.; Byrne, B.J. Long-term skeletal muscle protection after gene transfer in a mouse model of LGMD-2D. Mol. Ther. J. Am. Soc. Gene Ther. 2007, 15, 1775–1781. [Google Scholar] [CrossRef]
  316. Fougerousse, F.; Bartoli, M.; Poupiot, J.; Arandel, L.; Durand, M.; Guerchet, N.; Gicquel, E.; Danos, O.; Richard, I. Phenotypic Correction of α-Sarcoglycan Deficiency by Intra-arterial Injection of a Muscle-specific Serotype 1 rAAV Vector. Mol. Ther. J. Am. Soc. Gene Ther. 2007, 15, 53–61. [Google Scholar] [CrossRef]
  317. Cordier, L.; Hack, A.A.; Scott, M.O.; Barton-Davis, E.R.; Gao, G.; Wilson, J.M.; McNally, E.M.; Sweeney, H.L. Rescue of skeletal muscles of gamma-sarcoglycan-deficient mice with adeno-associated virus-mediated gene transfer. Mol. Ther. J. Am. Soc. Gene Ther. 2000, 1, 119–129. [Google Scholar] [CrossRef]
  318. Bartoli, M.; Roudaut, C.; Martin, S.; Fougerousse, F.; Suel, L.; Poupiot, J.; Gicquel, E.; Noulet, F.; Danos, O.; Richard, I. Safety and efficacy of AAV-mediated calpain 3 gene transfer in a mouse model of limb-girdle muscular dystrophy type 2A. Mol. Ther. J. Am. Soc. Gene Ther. 2006, 13, 250–259. [Google Scholar] [CrossRef]
  319. Duan, D. Systemic AAV Micro-dystrophin Gene Therapy for Duchenne Muscular Dystrophy. Mol. Ther. 2018, 26, 2337–2356. [Google Scholar] [CrossRef] [Green Version]
  320. Bowles, D.E.; McPhee, S.W.J.; Li, C.; Gray, S.J.; Samulski, J.J.; Camp, A.S.; Li, J.; Wang, B.; Monahan, P.E.; Rabinowitz, J.E.; et al. Phase 1 Gene Therapy for Duchenne Muscular Dystrophy Using a Translational Optimized AAV Vector. Mol. Ther. 2012, 20, 443–455. [Google Scholar] [CrossRef] [Green Version]
  321. World Health Organization. Cardiovascular Diseases (CVDs). Available online: https://www.who.int/en/news-room/fact-sheets/detail/cardiovascular-diseases-(cvds) (accessed on 1 December 2022).
  322. Jiang, H.; Lillicrap, D.; Patarroyo-White, S.; Liu, T.; Qian, X.; Scallan, C.D.; Powell, S.; Keller, T.; McMurray, M.; Labelle, A.; et al. Multiyear therapeutic benefit of AAV serotypes 2, 6, and 8 delivering factor VIII to hemophilia A mice and dogs. Blood 2006, 108, 107–115. [Google Scholar] [CrossRef] [Green Version]
  323. Zsebo, K.; Yaroshinsky, A.; Rudy, J.J.; Wagner, K.; Greenberg, B.; Jessup, M.; Hajjar, R.J. Long-Term Effects of AAV1/SERCA2a Gene Transfer in Patients With Severe Heart Failure. Circ. Res. 2014, 114, 101–108. [Google Scholar] [CrossRef]
  324. Nguyen, T.H.; Ferry, N. Liver gene therapy: Advances and hurdles. Gene Ther. 2004, 11, S76–S84. [Google Scholar] [CrossRef] [Green Version]
  325. Członkowska, A.; Litwin, T.; Dusek, P.; Ferenci, P.; Lutsenko, S.; Medici, V.; Rybakowski, J.K.; Weiss, K.H.; Schilsky, M.L. Wilson disease. Nat. Rev. Dis. Prim. 2018, 4, 21. [Google Scholar] [CrossRef]
  326. Murillo, O.; Collantes, M.; Gazquez, C.; Moreno, D.; Hernandez-Alcoceba, R.; Barberia, M.; Ecay, M.; Tamarit, B.; Douar, A.; Ferrer, V.; et al. High value of (64)Cu as a tool to evaluate the restoration of physiological copper excretion after gene therapy in Wilson’s disease. Mol. Ther. Methods Clin. Dev. 2022, 26, 98–106. [Google Scholar] [CrossRef]
  327. Murillo, O.; Moreno, D.; Gazquez, C.; Barberia, M.; Cenzano, I.; Navarro, I.; Uriarte, I.; Sebastian, V.; Arruebo, M.; Ferrer, V.; et al. Liver Expression of a MiniATP7B Gene Results in Long-Term Restoration of Copper Homeostasis in a Wilson Disease Model in Mice. Hepatology 2019, 70, 108–126. [Google Scholar] [CrossRef]
  328. Crigler, J.F., Jr.; Najjar, V.A. Congenital familial nonhemolytic jaundice with kernicterus. Pediatrics 1952, 10, 169–180. [Google Scholar]
  329. Gene Therapy for Severe Crigler Najjar Syndrome. Available online: https://ClinicalTrials.gov/show/NCT03466463 (accessed on 1 December 2022).
  330. Shi, X.; Aronson, S.J.; Ten Bloemendaal, L.; Duijst, S.; Bakker, R.S.; de Waart, D.R.; Bortolussi, G.; Collaud, F.; Oude Elferink, R.P.; Muro, A.F.; et al. Efficacy of AAV8-hUGT1A1 with Rapamycin in neonatal, suckling, and juvenile rats to model treatment in pediatric CNs patients. Mol. Ther. Methods Clin. Dev. 2021, 20, 287–297. [Google Scholar] [CrossRef] [PubMed]
  331. Golden, S.H.; Robinson, K.A.; Saldanha, I.; Anton, B.; Ladenson, P.W. Clinical review: Prevalence and incidence of endocrine and metabolic disorders in the United States: A comprehensive review. J. Clin. Endocrinol. Metab. 2009, 94, 1853–1878. [Google Scholar] [CrossRef] [PubMed]
  332. Wilson, J.D. Prospects for research for disorders of the endocrine system. JAMA 2001, 285, 624–627. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  333. Barzon, L.; Bonaguro, R.; Palù, G.; Boscaro, M. New perspectives for gene therapy in endocrinology. Eur. J. Endocrinol. 2000, 143, 447–466. [Google Scholar] [CrossRef] [Green Version]
  334. Chellappan, D.K.; Sivam, N.S.; Teoh, K.X.; Leong, W.P.; Fui, T.Z.; Chooi, K.; Khoo, N.; Yi, F.J.; Chellian, J.; Cheng, L.L.; et al. Gene therapy and type 1 diabetes mellitus. Biomed. Pharmacother. 2018, 108, 1188–1200. [Google Scholar] [CrossRef]
  335. Atkinson, M.A. Chapter 32—Type 1 Diabetes Mellitus. In Williams Textbook of Endocrinology, 3rd ed.; Melmed, S., Polonsky, K.S., Larsen, P.R., Kronenberg, H.M., Eds.; Elsevier: Philadelphia, PA, USA, 2016; pp. 1451–1483. [Google Scholar] [CrossRef]
  336. Flores, R.R.; Zhou, L.; Robbins, P.D. Expression of IL-2 in β cells by AAV8 gene transfer in pre-diabetic NOD mice prevents diabetes through activation of FoxP3-positive regulatory T cells. Gene Ther. 2014, 21, 715–722. [Google Scholar] [CrossRef]
  337. Zhang, Y.C.; Pileggi, A.; Agarwal, A.; Molano, R.D.; Powers, M.; Brusko, T.; Wasserfall, C.; Goudy, K.; Zahr, E.; Poggioli, R.; et al. Adeno-Associated Virus-Mediated IL-10 Gene Therapy Inhibits Diabetes Recurrence in Syngeneic Islet Cell Transplantation of NOD Mice. Diabetes 2003, 52, 708–716. [Google Scholar] [CrossRef] [Green Version]
  338. El Khatib, M.; Sakuma, T.; Tonne, J.M.; Mohamed, M.S.; Holditch, S.J.; Lu, B.; Kudva, Y.C.; Ikeda, Y. β-Cell-targeted blockage of PD1 and CTLA4 pathways prevents development of autoimmune diabetes and acute allogeneic islets rejection. Gene Ther. 2015, 22, 430–438. [Google Scholar] [CrossRef] [Green Version]
  339. RGX-314 Gene Therapy Administered in the Suprachoroidal Space for Participants With Diabetic Retinopathy (DR) without Center Involved-Diabetic Macular Edema (CI-DME). Available online: https://ClinicalTrials.gov/show/NCT04567550 (accessed on 1 December 2022).
  340. Borden, K.L. Pondering the promyelocytic leukemia protein (PML) puzzle: Possible functions for PML nuclear bodies. Mol. Cell. Biol. 2002, 22, 5259–5269. [Google Scholar] [CrossRef] [Green Version]
  341. Almaghrabi, S.; Azzouz, M.; Tazi Ahnini, R. AAV9-mediated AIRE gene delivery clears circulating antibodies and tissue T-cell infiltration in a mouse model of autoimmune polyglandular syndrome type-1. Clin. Transl. Immunol. 2020, 9, e1166. [Google Scholar] [CrossRef]
  342. Luo, J.; Luo, Y.; Sun, J.; Zhou, Y.; Zhang, Y.; Yang, X. Adeno-associated virus-mediated cancer gene therapy: Current status. Cancer Lett. 2015, 356, 347–356. [Google Scholar] [CrossRef] [Green Version]
  343. WHO. Cancer. Available online: https://www.who.int/news-room/fact-sheets/detail/cancer (accessed on 1 December 2022).
  344. Kanazawa, T.; Mizukami, H.; Okada, T.; Hanazono, Y.; Kume, A.; Nishino, H.; Takeuchi, K.; Kitamura, K.; Ichimura, K.; Ozawa, K. Suicide gene therapy using AAV-HSVtk/ganciclovir in combination with irradiation results in regression of human head and neck cancer xenografts in nude mice. Gene Ther. 2003, 10, 51–58. [Google Scholar] [CrossRef] [Green Version]
  345. Kim, J.Y.; Kim, J.H.; Khim, M.; Lee, H.S.; Jung, J.H.; Moon, D.H.; Jeong, S.; Lee, H. Persistent anti-tumor effects via recombinant adeno-associated virus encoding herpes thymidine kinase gene monitored by PET-imaging. Oncol. Rep. 2011, 25, 1263–1269. [Google Scholar] [CrossRef] [Green Version]
  346. Mizuno, M.; Yoshida, J.; Colosi, P.; Kurtzman, G. Adeno-associated Virus Vector Containing the Herpes Simplex Virus Thymidine Kinase Gene Causes Complete Regression of Intracerebrally Implanted Human Gliomas in Mice, in Conjunction with Ganciclovir Administration. Jpn. J. Cancer Res. 1998, 89, 76–80. [Google Scholar] [CrossRef]
  347. Kojima, Y.; Honda, K.; Hamada, H.; Kobayashi, N. Oncolytic gene therapy combined with double suicide genes for human bile duct cancer in nude mouse models. J. Surg. Res. 2009, 157, e63–e70. [Google Scholar] [CrossRef]
  348. Pan, J.G.; Zhou, X.; Luo, R.; Han, R.F. The adeno-associated virus-mediated HSV-TK/GCV suicide system: A potential strategy for the treatment of bladder carcinoma. Med. Oncol. 2012, 29, 1938–1947. [Google Scholar] [CrossRef]
  349. Borel, F.; Kay, M.A.; Mueller, C. Recombinant AAV as a platform for translating the therapeutic potential of RNA interference. Mol. Ther. J. Am. Soc. Gene Ther. 2014, 22, 692–701. [Google Scholar] [CrossRef] [Green Version]
  350. Sun, A.; Tang, J.; Terranova, P.F.; Zhang, X.; Thrasher, J.B.; Li, B. Adeno-associated virus-delivered short hairpin-structured RNA for androgen receptor gene silencing induces tumor eradication of prostate cancer xenografts in nude mice: A preclinical study. Int. J. Cancer 2010, 126, 764–774. [Google Scholar] [CrossRef]
  351. Kota, J.; Chivukula, R.R.; O’Donnell, K.A.; Wentzel, E.A.; Montgomery, C.L.; Hwang, H.W.; Chang, T.C.; Vivekanandan, P.; Torbenson, M.; Clark, K.R.; et al. Therapeutic microRNA delivery suppresses tumorigenesis in a murine liver cancer model. Cell 2009, 137, 1005–1017. [Google Scholar] [CrossRef] [Green Version]
  352. Lu, L.; Luo, S.T.; Shi, H.S.; Li, M.; Zhang, H.L.; He, S.S.; Liu, Y.; Pan, Y.; Yang, L. AAV2-mediated gene transfer of VEGF-Trap with potent suppression of primary breast tumor growth and spontaneous pulmonary metastases by long-term expression. Oncol. Rep. 2012, 28, 1332–1338. [Google Scholar] [CrossRef] [Green Version]
  353. Wu, Q.J.; Gong, C.Y.; Luo, S.T.; Zhang, D.M.; Zhang, S.; Shi, H.S.; Lu, L.; Yan, H.X.; He, S.S.; Li, D.D.; et al. AAV-mediated human PEDF inhibits tumor growth and metastasis in murine colorectal peritoneal carcinomatosis model. BMC Cancer 2012, 12, 129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  354. Tse, L.Y.; Sun, X.; Jiang, H.; Dong, X.; Fung, P.W.; Farzaneh, F.; Xu, R. Adeno-associated virus-mediated expression of kallistatin suppresses local and remote hepatocellular carcinomas. J. Gene Med. 2008, 10, 508–517. [Google Scholar] [CrossRef] [PubMed]
  355. Hunter, J.E.; Ramos, L.; Wolfe, J.H. Viral Vectors in the CNS. In Reference Module in Neuroscience and Biobehavioral Psychology; Elsevier: Amsterdam, The Netherlands, 2017. [Google Scholar] [CrossRef]
  356. Hermonat, P.L.; Quirk, J.G.; Bishop, B.M.; Han, L. The packaging capacity of adeno-associated virus (AAV) and the potential for wild-type-plus AAV gene therapy vectors. FEBS Lett. 1997, 407, 78–84. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  357. Flotte, T.R.; Carter, B.J. Adeno-Associated Viral Vectors. In Gene Therapy Technologies, Applications and Regulations; John Wiley & Sons Ltd.: Hoboken, NJ, USA, 1999; pp. 109–125. [Google Scholar] [CrossRef]
  358. Verdera, H.C.; Kuranda, K.; Mingozzi, F. AAV Vector Immunogenicity in Humans: A Long Journey to Successful Gene Transfer. Mol. Ther. J. Am. Soc. Gene Ther. 2020, 28, 723–746. [Google Scholar] [CrossRef]
  359. Ertl, H.C.J. Immunogenicity and toxicity of AAV gene therapy. Front. Immunol. 2022, 13, 975803. [Google Scholar] [CrossRef]
  360. Fitzpatrick, Z.; Leborgne, C.; Barbon, E.; Masat, E.; Ronzitti, G.; van Wittenberghe, L.; Vignaud, A.; Collaud, F.; Charles, S.; Simon Sola, M.; et al. Influence of Pre-existing Anti-capsid Neutralizing and Binding Antibodies on AAV Vector Transduction. Mol. Ther. Methods Clin. Dev. 2018, 9, 119–129. [Google Scholar] [CrossRef] [Green Version]
  361. Denard, J.; Marolleau, B.; Jenny, C.; Rao, T.N.; Fehling, H.J.; Voit, T.; Svinartchouk, F. C-reactive protein (CRP) is essential for efficient systemic transduction of recombinant adeno-associated virus vector 1 (rAAV-1) and rAAV-6 in mice. J. Virol. 2013, 87, 10784–10791. [Google Scholar] [CrossRef] [Green Version]
  362. Murphy, S.L.; Li, H.; Mingozzi, F.; Sabatino, D.E.; Hui, D.J.; Edmonson, S.A.; High, K.A. Diverse IgG subclass responses to adeno-associated virus infection and vector administration. J. Med. Virol. 2009, 81, 65–74. [Google Scholar] [CrossRef] [Green Version]
  363. Mingozzi, F.; Meulenberg, J.J.; Hui, D.J.; Basner-Tschakarjan, E.; Hasbrouck, N.C.; Edmonson, S.A.; Hutnick, N.A.; Betts, M.R.; Kastelein, J.J.; Stroes, E.S.; et al. AAV-1-mediated gene transfer to skeletal muscle in humans results in dose-dependent activation of capsid-specific T cells. Blood 2009, 114, 2077–2086. [Google Scholar] [CrossRef]
  364. Louis Jeune, V.; Joergensen, J.A.; Hajjar, R.J.; Weber, T. Pre-existing anti-adeno-associated virus antibodies as a challenge in AAV gene therapy. Hum. Gene Ther. Methods 2013, 24, 59–67. [Google Scholar] [CrossRef] [Green Version]
  365. Wang, L.; Calcedo, R.; Bell, P.; Lin, J.; Grant, R.L.; Siegel, D.L.; Wilson, J.M. Impact of pre-existing immunity on gene transfer to nonhuman primate liver with adeno-associated virus 8 vectors. Hum. Gene Ther. 2011, 22, 1389–1401. [Google Scholar] [CrossRef] [Green Version]
  366. Herzog, R.W.; Biswas, M. Neutralizing the Neutralizers in AAV Gene Therapy. Mol. Ther. 2020, 28, 1741–1742. [Google Scholar] [CrossRef]
  367. Zengel, J.; Carette, J.E. Structural and cellular biology of adeno-associated virus attachment and entry. Adv. Virus Res. 2020, 106, 39–84. [Google Scholar] [CrossRef]
  368. Chowdhury, E.A.; Meno-Tetang, G.; Chang, H.Y.; Wu, S.; Huang, H.W.; Jamier, T.; Chandran, J.; Shah, D.K. Current progress and limitations of AAV mediated delivery of protein therapeutic genes and the importance of developing quantitative pharmacokinetic/pharmacodynamic (PK/PD) models. Adv. Drug Deliv. Rev. 2021, 170, 214–237. [Google Scholar] [CrossRef]
  369. Kang, L.; Jin, S.; Wang, J.; Lv, Z.; Xin, C.; Tan, C.; Zhao, M.; Wang, L.; Liu, J. AAV vectors applied to the treatment of CNS disorders: Clinical status and challenges. J. Control. Release 2023, 355, 458–473. [Google Scholar] [CrossRef]
  370. Colle, M.A.; Piguet, F.; Bertrand, L.; Raoul, S.; Bieche, I.; Dubreil, L.; Sloothaak, D.; Bouquet, C.; Moullier, P.; Aubourg, P.; et al. Efficient intracerebral delivery of AAV5 vector encoding human ARSA in non-human primate. Hum. Mol. Genet. 2010, 19, 147–158. [Google Scholar] [CrossRef] [Green Version]
  371. Zerah, M.; Piguet, F.; Colle, M.A.; Raoul, S.; Deschamps, J.Y.; Deniaud, J.; Gautier, B.; Toulgoat, F.; Bieche, I.; Laurendeau, I.; et al. 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. 2015, 26, 113–124. [Google Scholar] [CrossRef]
  372. Zieger, M.; Borel, F.; Greer, C.; Gernoux, G.; Blackwood, M.; Flotte, T.R.; Mueller, C. Liver-directed SERPINA1 gene therapy attenuates progression of spontaneous and tobacco smoke-induced emphysema in α1-antitrypsin null mice. Mol. Ther. Methods Clin. Dev. 2022, 25, 425–438. [Google Scholar] [CrossRef]
Figure 1. Variant tropisms of AAV serotypes.
Figure 1. Variant tropisms of AAV serotypes.
Cells 12 00785 g001
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Issa, S.S.; Shaimardanova, A.A.; Solovyeva, V.V.; Rizvanov, A.A. Various AAV Serotypes and Their Applications in Gene Therapy: An Overview. Cells 2023, 12, 785. https://doi.org/10.3390/cells12050785

AMA Style

Issa SS, Shaimardanova AA, Solovyeva VV, Rizvanov AA. Various AAV Serotypes and Their Applications in Gene Therapy: An Overview. Cells. 2023; 12(5):785. https://doi.org/10.3390/cells12050785

Chicago/Turabian Style

Issa, Shaza S., Alisa A. Shaimardanova, Valeriya V. Solovyeva, and Albert A. Rizvanov. 2023. "Various AAV Serotypes and Their Applications in Gene Therapy: An Overview" Cells 12, no. 5: 785. https://doi.org/10.3390/cells12050785

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

Article Metrics

Back to TopTop