Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2013 Nov 1.
Published in final edited form as: Neurobiol Dis. 2011 Oct 7;48(2):179–188. doi: 10.1016/j.nbd.2011.09.014

Viral Vectors for Gene Delivery to the Central Nervous System

Thomas B Lentz 1, Steven J Gray 1, R Jude Samulski 1,*
PMCID: PMC3293995  NIHMSID: NIHMS336021  PMID: 22001604

Abstract

The potential benefits of gene therapy for neurological diseases such as Parkinson’s, Amyotrophic Lateral Sclerosis (ALS), Epilepsy, and Alzheimer’s are enormous. Even a delay in the onset of severe symptoms would be invaluable to patients suffering from these and other diseases. Significant effort has been placed in developing vectors capable of delivering therapeutic genes to the CNS in order to treat neurological disorders. At the forefront of potential vectors, viral systems have evolved to efficiently deliver their genetic material to a cell. The biology of different viruses offers unique solutions to the challenges of gene therapy, such as cell targeting, transgene expression and vector production. It is important to consider the natural biology of a vector when deciding whether it will be the most effective for a specific therapeutic function. In this review, we outline desired features of the ideal vector for gene delivery to the CNS and discuss how well available viral vectors compare to this model. Adeno-associated virus, retrovirus, adenovirus and herpesvirus vectors are covered. Focus is placed on features of the natural biology that have made these viruses effective tools for gene delivery with emphasis on their application in the CNS. Our goal is to provide insight into features of the optimal vector and which viral vectors can provide these features.

Keywords: Viral vectors, Central Nervous System, Gene Therapy, Adeno-associated virus, Retrovirus, Adenovirus, Herpesvirus

Introduction: Viruses as Vectors for Therapeutic Gene Delivery

The idea to treat human disease by delivering a therapeutic gene was first conceived as early as the 1970s (Friedmann and Roblin, 1972). The potential of viruses as vectors for gene delivery was recognized even then. Viruses are among the simplest biological agents capable of delivering genetic information to a cell and have evolved properties that make them useful as vectors for gene delivery (Shen and Post, 2007). Structural virions carry and protect viral genetic information (DNA and RNA), providing stability for transit throughout the human body. Determinants on the surface of the virion specify which cells the virus will enter while minimizing immunostimulatory potential within the host. Signals in the genetic material control gene expression from the virus in the infected cell. Finally, many viruses have developed mechanisms of replication that allow them to be propagated in cell culture. Though these innate properties have been advantageous in developing vectors for gene delivery, using viruses comes with potential risks as well. Many viruses cause disease and vectors derived from them have the potential to be pathogenic. The goal in designing vectors is to minimize this potential risk by using only components of the virus necessary for transduction of the target cell and expression of the transgene, while also preserving the efficiency of production. In this review, features of the ideal vector for delivering genes to the CNS are outlined. We then discuss the molecular biology of several viruses currently being developed as vectors and how they compare with our model of the ideal vector. Emphasis is placed on the application of these vectors in the CNS.

Model of the Ideal Vector for Gene Delivery to the CNS

Many viral systems have been developed as vectors for gene delivery and each offers unique solutions to the challenges of gene therapy. Determining which features a vector must possess to effectively treat a disease is the first step toward evaluating the particular strengths and weaknesses of different viral systems. To discuss the potential of different viral vectors for use in the CNS, we have developed a model of the ideal vector. One feature of the ideal vector is specific tropism for and highly efficient transduction of the target tissue and minimal transduction of ‘off-target’ cells, tissues or organs. Neurological diseases are often the result of loss or malfunction of a specific tissue within the CNS. For example, Parkinson’s is associated with the death of dopaminergic neurons in the substantia nigra. Treatment of diseases such as Parkinson’s requires that delivery of the therapeutic transgene be targeted to the affected tissue. Second, the ideal vector will express the transgene for a length of time and at a level as to have maximal therapeutic impact. It is important that these two parameters be empirically determined for the specific disease being treated. If the duration is too short or the level is too low, expression of the transgene may not produce a therapeutic effect. If too high, expression could have a cytotoxic effect on cells. High levels of expression may also be prone to elicit a host immune response against the therapeutic gene. A third feature of the ideal vector, and perhaps the most important, is that side effects such as vector related pathologies and/or host immune response be minimal. Some viruses (e.g. herpesviruses) lyse cells as a part of their normal replication cycle. For these viruses, the potential to generate replication competent particles during vector production is a significant concern. Other viruses (e.g. adenovirus) can be highly immunogenic. A strong host immune response can cause inflammation and/or mediate clearance of the transduced cells, resulting in lack of therapeutic effect and possibly consequences for patients’ health. Finally, methods of manufacture that are scalable and meet guidelines for therapeutic application must be available for the ideal vector. This is necessary to produce vector in high enough quantities and purity as to be practical for delivery to patients.

Comparison of the available viral vectors against the ideal vector reveals that the natural biology of the virus plays a significant role in whether a vector is suitable for use in the CNS. The packaging capacity of a virus can limit the size of transgene it is able to deliver. The viral proteins exposed on the surface of the virion can determine tissue tropism and immunogenicity. Even the form of the viral genome and how long it persists in a cell can govern the level and duration of transgene expression. However, innovations in the molecular biology of a vector can overcome the limitations imposed by the biology of the virus. Tissue tropism can be expanded and the specificity enhanced using methods of transencapsidation or pseudotyping to change receptor binding properties of a vector. Manipulation of the genetics, such as removal or introduction of functional sequences, can promote formation of more stable forms of a vector and thus increase the duration of transgene expression. Thus, it is important to consider the biology of the virus alongside the vectorology.

Adeno-Associated Virus Vectors

Adeno-associated virus (AAV) is a member of the family Parvoviridae in the genus Dependovirus (Berns and Parrish, 2007). Dependoviruses are non-autonomous, requiring co-infection with a helper virus (e.g. adenovirus or herpesvirus) in order to replicate, and non-pathogenic (Atchison et al., 1965; NIH, 2011). These aspects of its natural biology greatly reduce the risks of side effects associated with use of AAV for gene delivery. The virion is small, ~20 nm, consisting of a protein capsid that surrounds the viral genome (Cassinotti et al., 1988; Rose et al., 1971; Xie et al., 2002). The capsid is made up of viral capsid (Cap) proteins VP1, VP2, and VP3. These proteins facilitate binding of the virion to the cell surface receptor, which varies between serotypes. AAV2, the predominant serotype capsid of use in clinical applications, uses heparin sulfate as its primary receptor (Summerford and Samulski, 1998). Other serotypes have been shown to use sialic acid (Serotypes 1, 4, 5 and 6 (Kaludov et al., 2001; Wu et al., 2006b)), Laminin receptor (serotype 8 (Akache et al., 2006)), and N-linked galactose (serotype 9 (Shen et al., 2011)). The target tissue of natural infection has not been determined, but tropism can vary considerably between capsids of different serotypes (Chao et al., 2000; Rabinowitz et al., 2002; Wu et al., 2006a; Xiao et al., 1999). The genome of AAV is a single-stranded (ss) DNA of ~4.7 kb flanked on 5′ and 3′ ends by an inverted terminal repeat (ITR). The ITR is predicted to form a “T” shaped stem loop structure due to complementarity within its sequence, and functions as an origin for genome replication, which proceeds through a rolling hairpin mechanism (Berns and Parrish, 2007; Lusby et al., 1980; Nash et al., 2008; Straus et al., 1976; Tattersall and Ward, 1976). Second-strand synthesis occurs within the transduced cell and requires several cellular enzymes (Nash et al., 2008). Expression of the viral replication (Rep) proteins, 78, 68, 52, 40 (name reflects molecular weight), is also necessary for replication and packaging of the genome (Im and Muzyczka, 1990; Im and Muzyczka, 1992; Snyder et al., 1990). In the absence of helper virus the viral genome persists episomally in a latent state as monomers and multimers in linear and circular forms (Duan et al., 1998; Schnepp et al., 2005). Integration of the genome is observed at a relatively low frequency and has preference for a specific site on chromosome 19 (McCarty et al., 2004). However, integration is greatly reduced and loses site specificity in the absence of Rep 78/68 (as is the case with vectors derived from AAV).

Several different types of vectors have been derived from AAV, each using a different strategy to facilitate transgene expression. One requirement of all AAV vectors is that the transgene cassette be flanked by the ITR sequence. This is necessary for packaging the transgene into the viral capsid. AAV serotype 2 was the first serotype to be used extensively and, as a result, the AAV2 ITR is the most widely used. Another requirement is that the vector not exceed ~4.7 kb in size, the packaging capacity of the viral capsid (Dong et al., 2010; Dong et al., 1996; Hirsch et al., 2010; Lai et al., 2010; Wu et al., 2010). This packaging capacity is rather stringent, as vectors that exceed this size are packaged in truncated forms. Single-stranded AAV (ssAAV) vectors were the first to be developed and consist of a transgene expression cassette flanked on 5′ and 3′ ends by the ITR sequence (Samulski et al., 1989). In these vectors the transgene cassette simply replaces the coding region of wild-type (WT) AAV and is packaged as single-stranded DNA, as the name indicates. No viral genes are encoded by the vector, reducing the potential for host immune recognition of transduced cells. One limitation of these vectors is that transgene expression depends on second-strand synthesis, which is a rate limiting step in cell transduction (Ferrari et al., 1996). Self-complementary (scAAV) vectors were developed to overcome the dependence on second-strand synthesis (McCarty et al., 2001). scAAV vectors contain a mutated ITR positioned centrally between two copies of the transgene sequence, one inverted and complementary to the other. This allows the genome to self-anneal to form a double-stranded molecule. These vectors express the transgene earlier and to greater levels than ssAAV vectors, but limit the size of the transgene cassette to ~2.2 kb. However, limited coding capacity has prevented the use of scAAV and even ssAAV vectors for applications requiring large transgene cassettes, >5 kb. To circumvent this, dual vector approaches have been developed that can increase transgene capacity up to ~8 kb (Duan et al., 2001; Hirsch et al., 2010; Hirsch et al., 2009). In these strategies, the transgene cassette is separated over two vectors and assembly of the full-length transgene relies on co-infection of a single cell and subsequent recombination. Split (trans-splicing) vectors are designed with a splice donor site on the vector encoding the 5′ half of the transgene and a splice acceptor site on the vector encoding the 3′ half of the transgene such that recombination between the two occurs within an intron (Duan et al., 2000; Ghosh et al., 2011; Nakai et al., 2000; Sun et al., 2000; Yan et al., 2000). This approach enables expression of larger transgene cassettes, but the low efficiency of co-transduction of a single cell and vector recombination significantly reduces the level of transgene expression relative to ssAAV and scAAV vectors.

Several features of AAV vectors stack up well against our model of the ideal vector for gene delivery to the CNS. Though not thought to be a natural site of infection, AAV vectors transduce many different cell types of the CNS, depending on which capsid is used. This includes neurons, which most serotypes transduce with high efficiency, as well as astrocytes, glial and ependymal cells, which are selectively transduced well by a subset of capsids (Alisky et al., 2000; Broekman et al., 2006; Burger et al., 2004; Davidson et al., 2000; Kaplitt et al., 1994; Lawlor et al., 2009; Van der Perren et al., 2011; Van Vliet et al., 2008; Wang et al., 2003). The capsid of serotype 9 has even been shown to cross the blood brain barrier (Duque et al., 2009; Foust and Kaspar, 2009; Foust et al., 2009; Gray et al., 2011). AAV vectors can also engender long-term, stable transgene expression in the CNS, persisting in the murine brain for >6 months (Klein et al., 1999). Other reports have demonstrated persistence in other tissues for >6 years in primates and >8 years in dogs (Niemeyer et al., 2009; Rivera et al., 2005; Stieger et al., 2009). No significant side effects have been found associated with delivery of AAV vectors (Bessis et al., 2004). The possibility that AAV vectors may be correlated with an increased risk of tumorogenesis has been presented, but a study covering over 600 animals found no association (Bell et al., 2006; Bell et al., 2005; Kay, 2007). Finally, scalable systems for vector production have been developed and are currently in use (Grieger et al., 2006; Grieger, personal communication; Lock et al., 2010; Vandenberghe et al., 2010).

While AAV vectors are being tested for clinical use in the CNS, the field is also moving quickly to improve upon these vectors. Vectors derived from AAV, particularly serotype 2, are the predominant viral vectors of use to treat neurological diseases, accounting for 76% of therapy trials in the CNS (Wiley, 2011). Trials targeting Parkinson’s, Amyotrophic Lateral Sclerosis (ALS), Alzheimer’s, Batten Disease, and Epilepsy are ongoing. Strategies are being developed that may expand this list to include Huntington’s Disease, Mucopolysaccharidoses (MPS) and Spinocerebellar Ataxia (McCown, 2011). One focus is to expand the list of possible tissue targets by synthesizing vectors with novel tropism and increased efficiency/specificity of transduction (Michelfelder and Trepel, 2009; Muzyczka and Warrington, 2005; Van Vliet et al., 2008). Historically, AAV2 has been used more widely than other serotypes and early vectors employed the ITR of serotype 2 which could only be packaged in the corresponding capsid. Development of a cross-packaging system allows a single vector carrying the serotype 2 ITR to be trans-encapsidated with the capsids of most AAV serotypes (Rabinowitz et al., 2002). This system provides a way to customize the capsid of a vector for tissue specificity for the CNS and can overcome the possible complication associated with pre-existing immunity against AAV2 (Taymans et al., 2007; Van der Perren et al., 2011; Vandenberghe et al., 2006). Random diversification and selection of capsids in a process termed molecular evolution has further increased the potential for customization. This method has yielded capsids with improved transduction of astrocytes in vitro, ~15-fold, and in vivo, ~6-fold, when injected into the rat striatum (Koerber et al., 2009). Similar strategies have been employed to customize vectors for specific therapeutic applications or disease states. For example, shuffling the Cap protein coding sequence of different serotypes has produced novel capsids capable of crossing the seizure-compromised blood brain barrier in a rat model for epilepsy (Gray et al., 2010a). Additionally, Chen et al. cycled a phage display library of random peptides in normal, LINCL, or MPS VII mice, injecting the library into the tail vein and recovering it from the brain (Chen et al., 2009b). The recovered peptides were then incorporated into an AAV2 capsid, which specifically targeted to the cerebral vascular endothelial cells after i.v. injection in mice. The peptides identified were highly specific for the disease state they were selected in, and were not cross-compatible with WT mice. Another focus of vector engineering is to reduce the potential risk of vector mobilization. Evidence that vectors may be replicated in the presence of WT AAV and helper-virus infection has been produced from experiments in cell culture and in vivo (Afione et al., 1996; Cheung et al., 1980). To address this, novel ITRs have been developed that are not replicated by Rep proteins of WT AAV and thus, cannot be mobilized (Hewitt et al., 2009; Hewitt and Samulski, 2010).

In summary, AAV vectors are currently the preferred gene delivery vehicle for CNS applications. They provide efficient gene transfer, long-term transgene expression, minimal pathogenicity, low immunogenicity, and scalable manufacture for clinical applications. However, AAV vectors are hindered by their 4.7-kb packaging limit. For a detailed review of the many applications of AAV in the CNS, see references (Gray et al., 2010b; McCown, 2011).

Retrovirus vectors

Retroviridae is a family of single-stranded (ss) RNA viruses that reverse transcribe their genetic material into DNA and integrate into the host genome as part of their life cycle (Goff, 2007). Genera of this family can be categorized as simple (alpharetroviruses, betaretroviruses, gammaretroviruses, and epsilonretroviruses) or complex (deltaretroviruses, lentiviruses and spumaviruses) retroviruses depending on the proteins they encode. Both simple and complex retroviruses have Gag, Pol and Env genes (Leis et al., 1988). In addition, complex retroviruses encode several small regulatory proteins that play a role in virus-host interaction. Retroviral virions average ~100 nm in size and are composed of a phospholipid envelope surrounding a capsid, within which two copies of the ssRNA genome, both positive sense, and several viral proteins are organized into a nucleocapsid (Chen et al., 2009a; Goff, 2007; Liu et al., 2010). The envelope is acquired upon budding from the host cell plasma membrane and is comprised of a phospholipid bilayer containing several viral glycoproteins, encoded by the Env gene (Forster et al., 2005; Johnson, 2011). Envelope glycoprotein SU (surface) interacts with the cell surface receptor and TM (transmembrane) appears to play a role in membrane fusion. A variety of cell surface receptors are used by different retroviruses, and target predominantly T cells and macrophages (Miller, 1996; Sommerfelt, 1999; Weiss and Tailor, 1995). Inside the envelope the Gag poly-protein is the major structural component, interacting with both the interior surface of the envelope membrane as well as the RNA genome (Freed, 1998; Scarlata and Carter, 2003). When the virus buds from a cell, Gag is cleaved by the viral protease, encoded by the Pol gene, into matrix (MA), capsid (CA), and nucleocapsid (NC) subunits. A structural rearrangement of the virion, termed maturation, then takes place, forming the conical capsid and nucleocapsid, which are referred to collectively as the viral core (Briggs et al., 2003; de Marco et al., 2010). Reverse transcription of the ssRNA by the viral reverse transcriptase, encoded by the Pol gene, takes place within the core, during or shortly before entering a cell. This process results in the synthesis of a linear double-stranded (ds) DNA genome and formation of a pre-integration complex (Telesnitsky and Goff, 1997). This process is primed by tRNA at the primer binding site (PBS) and requires several template switches, which are facilitated by long terminal repeat (LTR) sequences present on the termini of the genome. The pre-integration complex then enters the nucleus either through active transport (complex retroviruses) or upon cell division (simple retroviruses) depending on the virus (Bukrinsky et al., 1993; Lewis and Emerman, 1994). Once in the nucleus, the viral integrase, encoded by the Pol gene, facilitates integration of the linear dsDNA into the genome of the cell, forming a provirus which serves as the template for viral gene expression (Brown et al., 1989).

Vectors for gene delivery have been generated from simple (e.g. Moloney murine leukemia virus; gammaretrovirus) as well as complex (e.g. HIV-1; lentivirus) retroviruses (Escors and Breckpot, 2010; Maier et al., 2010). The predominant difference between vectors derived from these two types of retroviruses is in their reliance on cell division for transduction. Vectors derived from simple retroviruses require passage through the cell cycle for transduction while vectors derived from complex retroviruses are capable of transducing non-dividing cells (Lewis and Emerman, 1994). This represents a distinct advantage of complex over simple retroviral vectors for gene delivery to the quiescent cells of the CNS (Escors and Breckpot, 2010; Jakobsson and Lundberg, 2006; Wong et al., 2006). Though simple retroviruses have been employed to study aspects of neurogenesis, lentiviral vectors have become the predominant retroviral vector for gene delivery to the CNS.

Vectors derived from lentivirus have a majority of the viral genes deleted and retain only the cis-acting sequence elements necessary for nuclear export of the RNA, RNA dimerization, packaging, and reverse transcription (Dull et al., 1998; Zufferey et al., 1997). This yields vectors with an available packaging capacity of ~8 kb. Comprehensive analysis of lentiviral vector packaging capacity revealed that this is not an absolute limit, but rather that there is a near-linear negative correlation between increasing vector size and virus production (Kumar et al., 2001). Packaging cell lines have been developed for simple as well as complex retroviral systems that provide the viral genes necessary for vector production in trans (Cockrell et al., 2006; Danos and Mulligan, 1988; Kafri et al., 1999; Klages et al., 2000; Markowitz et al., 1988). Modern vectors have been modified in an effort to eliminate integration into the host genome. Initially seen as an advantage, integration comes with a significant risk of insertional mutagenesis, as was realized with the incidence of leukemia in 4 of 20 patients treated for X-linked SCID (Bokhoven et al., 2009; Hacein-Bey-Abina et al., 2003; Pike-Overzet et al., 2007). It is not clear if the risk of insertional mutagenesis in the ex vivo cell therapy approach used in this X-linked SCID trial reflects that of in vivo treatment of non-dividing cells, such as neurons of the CNS, but this risk has hindered the progress of these vectors. One strategy to overcome this challenge has been to try directing integration to ‘safe’, heterochromatin regions of the genome, which may minimize the risk of gene activation (Gijsbers et al., 2010). Alternatively, introduction of self-inactivating (SIN) mutations, which knock out promoter activity of the LTR, has greatly reduced the risk of insertional gene activation (Miyoshi et al., 1998; Zufferey et al., 1998). Another major advancement was the development of non-integrating lentiviral (NIL) vectors, which carry either mutant integrase or mutations in their LTRs that inhibit integrase binding (Apolonia et al., 2007; Philippe et al., 2006; Sarkis et al., 2008). Though still observed at ~0.35 – 2.3 %, integration is greatly reduced with NIL vectors (Cornu and Cathomen, 2007). Instead of integrating the vector genome persists episomally in linear and circular forms (Cara and Reitz, 1997; Philpott and Thrasher, 2007). Efficiency of transduction of target cells, including neurons, is not significantly affected by these mutations and is similar to that of integrating vectors (Cornu and Cathomen, 2007; Rahim et al., 2009; Sarkis et al., 2008). Recently, removal of a sequence element involved in plus-strand DNA synthesis was shown to further reduce integration and increase the efficiency of formation of circular episomes (Kantor et al., 2011).

Many features of lentiviral vectors compare well with our model of the ideal vector. Though these vectors do not naturally infect cells of the CNS, there is significant potential for modifying tissue tropism for specific therapeutic applications. The mechanism for envelopment of lentiviruses allows vectors to be pseudotyped with envelope proteins from other viruses (e.g. vesicular stomatitis virus G) (Bischof and Cornetta, 2010; Chen et al., 1996; Naldini et al., 1996). Taking advantage of this feature of the biology of the virus has generated vectors capable of transducing neurons and astroglial cells (Colin et al., 2009; Desmaris et al., 2001; Greenberg et al., 2007; Jakobsson et al., 2006; Rahim et al., 2009). Also, innovations in the genetics of the vector, leading to the creation of SIN and NIL vectors, have reduced the risk of integration and vector-related pathologies. Finally, systems of vector production have been developed that can be scaled up for manufacture of clinical reagents (Kafri et al., 1999; Klages et al., 2000). However, one shortcoming of NIL vectors is that duration and level of transgene expression are not well characterized. Transgene expression has been shown to persist up to 3 months in the rat brain and as long as 9 months in other tissues, but more extensive studies are required to determine the limit of expression (Bayer et al., 2008; Yanez-Munoz et al., 2006).

Retroviral vectors are currently the second most commonly used viral vector for gene delivery to the CNS. Ex vivo transduction of cells with a Moloney leukemia virus vector expressing nerve growth factor (NGF) and introduction to the brain has been used in treatment of Alzheimer’s disease with moderate success (Tuszynski et al., 2005). A trial using a lentiviral vector delivering three genes involved in dopamine biosynthesis, ProSavin, to treat Parkinson’s disease is also underway (Oxford_Biomedica, 2011; Wiley, 2011). Preliminary studies in animal models suggest these vectors might also be used in treating Huntington and lysosomal storage diseases (Jakobsson and Lundberg, 2006; Lundberg et al., 2008; Wong et al., 2006).

NIL vectors are at the forefront of retroviral vector development and hold promise as tools for gene delivery to the CNS. Further reduction in the rate of integration and characterization of the longevity of transgene expression are necessary to meet our requirements for the ideal vector. For a review of the application of retroviral and/or lentiviral vectors in the CNS, see references (Jakobsson and Lundberg, 2006; Lundberg et al., 2008; Wong et al., 2006).

Adenovirus vectors

Adenoviruses belong to the family Adenoviridae and range in size from ~70 – 100 nm (Berk, 2007). The virion consists of an icosahedral protein capsid, composed of subunits hexon, penton and fiber, surrounding a core, which contains the genomic material of the virus (Reddy et al., 2010). Hexon is the predominant subunit of the capsid, making up the surface of the structure, while penton and fiber subunits reside only at the vertices, the latter protruding from the surface in a characteristic spike. Virus attachment to cells is mediated via interaction of the fiber protein with a specific cell surface receptor, the coxsackievirus and adenovirus receptor (CAR) (Bergelson et al., 1997; Tomko et al., 1997). This receptor is present on the surface of a number of tissues, including cells of the CNS. The adenoviral genome is a linear, dsDNA molecule of ~36 kb (Davison et al., 2003). Inverted terminal repeat (ITR) sequences flank the genome and serve as the origin for replication. Additionally, a cis element has been identified near the end of the genome that is required for efficient genome packaging (Hearing et al., 1987). Adenoviruses do not integrate into the host genome as part of their life-cycle and the frequency of integration events is very low. Instead, the genome is maintained as a linear episome in the nucleus of a cell (Hillgenberg et al., 2001).

Adenoviral vectors have gone through several generational steps of development. First generation vectors contain deletions of the E1 and/or E3 regions of the genome, the former encodes proteins necessary for early gene expression the latter is dispensable for replication and packaging (Bett et al., 1994). This yielded a transgene capacity of up to ~8.3 kb, but this generation of vectors retains a majority of the viral genes and cytoxicity related to expression of the viral genes has limited their utility. Second generation vectors have more extensive deletions of the adenoviral sequence, removing the E2 or E4 regions along with E1/3 (Amalfitano et al., 1998; Armentano et al., 1995). These vectors are incapable of replicating on their own and require helper vectors or cell lines to complement the function of these two regions. Removal of these additional regions has helped reduce inflammation and host immune response to the vectors, but it remains a problem. More recently, helper-dependent (‘gutted’) vectors were created that are devoid of most of the viral genes (Parks et al., 1996; Volpers and Kochanek, 2004). Helper-dependent vectors contain the minimal sequence elements necessary for production, the ITRs and the packaging signal, and have a capacity of ~36 kb (Parks, 2000). Similar to WT adenovirus, adenoviral vectors are maintained episomally as a linear molecule in the nucleus of a cell (Jager and Ehrhardt, 2009; Rauschhuber et al., 2011). Circular molecules can also be produced, but the appearance of these forms may depend on the sequence contained within the vector and/or the presence of adenovirus proteins (Kreppel and Kochanek, 2004). Integration of vector into the host genome is observed only minimally, even when extensive homology exists between the recombinant vector and the host genome (Stephen et al., 2010; Stephen et al., 2008).

Adenoviral vectors exhibit some of the features desired in our model of the ideal vector, but still have to overcome the risk of host immune response and the lack of helper-free production methods before they can be used extensively in the CNS. Helper-dependent vectors have been shown to transduce of neuronal, astroglial, and human glioma cells, demonstrating that targeting of neuronal tissues is possible with these vectors (Candolfi et al., 2006; Candolfi et al., 2007). Additionally, methods of modifying vectors to broaden tropism and reduce off-target transduction have been developed (Everts and Curiel, 2004; Glasgow et al., 2006). Transductional targeting is perhaps the most promising of these methods and employs bi-specific molecules that bind to the fiber protein, concealing the ligand for CAR, and expose a new receptor binding ligand. Specificity for various tissues has been accomplished in this manner. Duration of transgene expression also shows promise for therapeutic applications. Expression from helper-dependent vectors persists significantly longer than earlier generation vectors, lasting >2 years in the liver of baboons (Morral et al., 1999; Morsy et al., 1998). The foremost challenge confronting adenovirus-derived vectors has been to eliminate host immune response to the vector (Bessis et al., 2004). Early generation vectors cause significant toxicity in the CNS as well as other tissues, resulting in clearance of transduced cells (Lozier et al., 2002; Thomas et al., 2001; Thomas et al., 2002). This has led to successful application of these vectors to treat cancers of the CNS, but has limited progress forward in treating other diseases (Germano et al., 2003; Immonen et al., 2004). Host immune response against helper-dependent vectors is reduced compared with that of earlier generation vectors, but remains a problem (Chirmule et al., 1999; Morral et al., 1999; Morsy et al., 1998). Development of scalable methods of manufacturing helper-free virus has not yet been achieved and is another issue that must be overcome. Production of adenoviral vectors require that viral genes, deleted in the vector sequence, be provided in trans from a complementing plasmid or cell line. Recombination between the complementing sequence and the vector sequence can produce vector that carries viral genes. One future direction of the field has been to develop systems of production that eliminate this risk. In a recent advance, Suzuki et al. have taken advantage of the natural packaging limit of adenoviral capsid to ensure that if helper virus is produced by recombination between vector and helper genomes it will not be packaged (Suzuki et al., 2011).

Adenovirus vectors have not yet achieved the success of AAV- and lentivirus-derived vectors for gene delivery to the CNS. Transgene coding capacity of ~36 kb and the ability to transduce cells types of the CNS are promising attributes. However, toxicity and immunogenicity currently impede the advancement of these vectors to the clinic.

Herpesvirus vectors

Herpesviridae is a broad family of viruses grouped by common biology and virion morphology (Pellett and Roizman, 2007). Vectors derived from herpes simplex virus (HSV) have come to dominate the field, but several members of this family have been considered for gene delivery (e.g. Epstein-Barr virus and Herpesvirus saimiri) (Griffiths et al., 2006; Robertson et al., 1996). For a detailed description of HSV and its life cycle, see reference (Roizman et al., 2007). The virion of HSV is large relative to other viruses, ~ 186 nm, and consists of a lipid envelope surrounding a capsid and a pleiomorphic layer called the tegument (Grunewald et al., 2003). The envelope is a phospholipid bilayer embedded with several viral glycoproteins and is acquired upon budding into membranous compartments in the cytoplasm of a cell. The tegument, found just beneath the envelope, is largely unstructured and contains at least 20 viral proteins and possibly viral and cellular RNA. The capsid has a regular icosahedral structure and is composed of capsomer subunits. Within the capsid, the linear dsDNA genome, ~152 kb, is associated with several viral proteins to make up the viral core. The genome can be described as having two distinct coding regions, unique long (UL) and unique short (US) coding regions, both flanked by terminal repeat sequences (Wadsworth et al., 1975). HSV encodes ~90 genes, the functions of which are numerous. However, many of these genes are dispensable and can be removed without inhibiting genome replication or packaging of the virus.

Infection with WT HSV can be either lytic or latent. Glycoproteins on the exterior of the envelope mediate entry of a cell by first binding glycosaminoglycans (i.e. heparan sulfate proteoglycan) on the cell surface (Shieh et al., 1992). Subsequent internalization steps involve additional interaction with one of several possible secondary receptors. Though the primary receptor can be found on many different cell types throughout the body, infection with HSV typically occurs at the cutaneous or mucosal epithelium where replication of the virus is initially lytic. The virus can also invade axons of sensory neurons in the affected area and undergo retrograde transport, up the axon, to the dorsal root ganglia (Bearer et al., 2000). Proteins in the tegument have specific roles upon infection of a cell, including involvement in travel up the axon. In fact, it has become apparent that one function of this layer of the virion is simply to transport these proteins to the infected cell, pre-assembled. Once the virus has reached the nucleus of a cell, the linear genome circularizes and is maintained as an episome with minimal integration (Garber et al., 1993; Strang and Stow, 2005). The genome can be maintained in this state for the life of the infected individual. After circularization, genome replication proceeds through one of two mechanisms, initially following a theta mode and later switching to a rolling-circle mode of synthesis (Boehmer and Lehman, 1997).

Several different types of vectors have been derived from HSV (Marconi et al., 2009; Shen and Post, 2007). Attenuated (‘replication-conditional’) vectors are deficient in expression of viral genes essential for replication in non-dividing, but not dividing cells (e.g. thymidine kinase and ribonucleotide reductase) (Wu et al., 1996). Because replication of these vectors is highly toxic to infected cells, they can be used to target proliferating cells for killing. Though not desirable for delivering therapeutic genes, these vectors have been used in the CNS, primarily for treatment of cancers such as glioblastoma multiforme (Andreansky et al., 1997; Lou, 2003; Markert et al., 2009; Mineta et al., 1995). More extensive removal of viral genes led to the creation of replication-defective vectors, which lack the ability to replicate autonomously and require complementing helper function for propagation (Krisky et al., 1998a; Marconi et al., 1996; Samaniego et al., 1998). Removal of immediate-early genes, necessary for expression of early and late genes, from these vectors greatly reduced vector toxicity. This eliminated cytotoxic vector replication within the cell and reduced host immune response to viral gene products. As a result, these vectors can deliver and facilitate expression of a transgene, which can persist for as long as ~3 weeks in cell culture and ~1 month in vivo (Krisky et al., 1998b; Lilley et al., 2001). Identification of the minimal cis-acting sequences necessary for virus replication, the cleavage/packaging signal and the viral origin, provided another significant advancement in the design of HSV vectors (Spaete and Frenkel, 1982). Amplicon vectors contain only these elements and have a packaging capacity of ~150 kb (Epstein, 2009; Marconi et al., 2009). These vectors have several advantages over attenuated or replication-defective vectors, including lack of viral genes that might cause cytotoxicity and maximal transgene coding capacity. Additionally, in instances in which the vector is significantly smaller than the packaging capacity of the virion, head-to-tail concatemers of the vector can be packaged (Saeki et al., 2003). The rolling circle mechanism of genome replication produces head-to-tail concatemers of the genome that are normally cleaved to liberate monomers for packaging. If these concatemers do not exceed capacity they can be packaged without cleavage. Delivering multiple copies of the transgene in each vector may produce increased levels of transgene expression in the transduced cells, another potential advantage. Transgene expression with these vectors has been demonstrated to persist for up to 7 months in the rat brain, but may not be stable over this period of time (Sun et al., 2003; Zhang et al., 2000). Due to the natural biology of the virus, all vectors derived from HSV are neurotropic and can efficiently transduce neurons of the CNS (Pakzaban et al., 1994; Wolfe et al., 1992). The envelope can be pseudotyped with the envelope proteins of other viruses to alter tissue tropism and reduce off-target transduction as well, as has been demonstrated with VSV-G protein (Anderson et al., 2000).

Possibly the greatest hurdle to the progress of HSV vectors has been the lack of a method of production that eliminates the potential for contaminating helper virus and yields high-titer vector. Production of HSV vectors requires complementation with helper vectors or cell lines that provide the necessary viral proteins in trans. Early versions of helper vectors could be packaged and were present at high levels in vector preparations. An improvement in production was the separation of helper vector sequence onto different cosmids, none of which carry the viral packaging signal (Fraefel et al., 1996). This system was able to produce vector with reduced levels of contaminating helper virus, but cytotoxic helper virus replication was not eliminated. A more recent method of production uses bacterial artificial chromosome (BAC) to express the complementing viral genes (Saeki et al., 2003). Working with a single BAC has technical advantages over working with multiple different cosmids. However, contamination with cytotoxic helper virus remains a problem. Further advancement in this area is needed in order to identify scalable and clinically relevant methods of vector manufacture.

HSV-derived vectors have several positive attributes but do not stand up well against our model of the ideal vector for gene delivery to the CNS. These vectors are able to be targeted for cells of the CNS and thus, possess one feature of our model for the ideal vector. It is also worth noting, that these vectors have the largest packaging capacity of all of the vectors we have discussed in this review. Though not a requirement of the ideal vector, this may serve as an advantage over other vector systems in certain applications. The shortcomings of these vectors are in the duration of transgene expression, the potential for harmful side effects associated with co-production of helper virus, and the lack of a scalable system for vector production.

HSV vectors have not yet been used in the clinic for therapeutic gene delivery to the CNS, but progress has been made in developing vectors for the treatment of neurological diseases. Amplicon vectors expressing the anti-apoptotic peptide Bcl-2, neurotrophic factors GDNF and BDNF, and enzymes of the dopamine biosynthesis pathway have shown protection against neuron degeneration in a rat model of Parkinson’s disease (Sun et al., 2005; Sun et al., 2003; Yamada et al., 1999). Also, replication-conditional vectors have been researched extensively for their oncolytic properties and applied in the clinic to treat glioblastoma multiforme and other cancers (Marconi et al., 2009; Markert et al., 2009; Markert et al., 2000; Shen and Post, 2007).

In summary, herpesvirus-derived vectors hold promise as tools for gene delivery to the CNS, but are still in their infancy. Further progress in the duration of transgene expression and scalable methods of safe vector production is required to compete with other available vector systems.

Conclusion

Several viral systems have been developed as vectors for gene delivery. We have described features of the ideal vector for clinical application in the CNS. Of the viral systems available, AAV vectors compare well against the ideal vector. AAV is minimally toxic and does not elicit a strong host immune response. Transduction can be achieved in several tissues of the CNS and therapeutic levels and duration of transgene expression have already been demonstrated. Additionally, scalable methods of production are readily available. Retroviral vectors have several promising attributes as well, including a larger packaging size, capacity for cell targeting and scalable production. However, the potential for vector integration and insertional mutagenesis has limited the use of these vectors. Continued improvement of non-integrating derivatives may eventually eliminate this risk and produce a vector with a stronger safety profile. Other viral systems, adenovirus and herpesvirus, still have several hurdles to overcome in meeting the standards of an ideal vector. Both are associated with potential pathologies related to strong host immune response (adenovirus) or cytotoxicity (herpesvirus). Factors such as lack of scalable production (for herpesvirus) and duration and level of transgene expression have also impeded the progress of these vectors. For these reasons adenovirus and herpesvirus vectors are currently not considered for most CNS applications. Thus, in most circumstances AAV is currently the most appropriate vector for therapeutic gene delivery to the CNS.

Figure 1.

Figure 1

Viral Vectors for Gene Delivery to the CNS

Table 1.

Comparison of viral vectors to the model of the ideal vector for gene delivery to the CNS.

Features of Ideal Vector → Transduction of target tissue Duration of transgene expression Minimal Pathogenesis Scalable production
AAV vectors +++ (neurons, astrocytes, glial and ependymal cells) +++ (6 months in brain, 6 years in other tissues of primates) +++ (no associated pathologies) +++ (large scale production of highly pure vector)
Retrovirus Vectors (NIL/Integrating) ++ (neurons and astroglial cells) +/+++ (3 months in brain, 9 months in other tissues of murines) +++/+(potential pathologies associated with integration) ++ (moderately scalable production of highly pure vector)
Adenovirus vectors ++ (neural, astroglial, and glioma cells) ++ (2 yrs in non-brain tissues of primates) − (immune response to vector and helper-virus contamination present pathologies) ++ (large scale production; helper-virus contamination)
HSV vectors + (neurons) + (7 months in brain of murine, unstable) − (helper-virus contamination presents pathologies) − (scalable production not yet achieved)

How well each vector compares with the ideal vector on each feature is indicated with ‘+’ or ‘−’ signs.

Acknowledgments

The authors would like to thank Adam Cockrell for critical review of the manuscript.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  1. Afione SA, et al. In vivo model of adeno-associated virus vector persistence and rescue. J Virol. 1996;70:3235–41. doi: 10.1128/jvi.70.5.3235-3241.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Akache B, et al. The 37/67-kilodalton laminin receptor is a receptor for adeno-associated virus serotypes 8, 2, 3, and 9. J Virol. 2006;80:9831–6. doi: 10.1128/JVI.00878-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Alisky JM, et al. Transduction of murine cerebellar neurons with recombinant FIV and AAV5 vectors. Neuroreport. 2000;11:2669–73. doi: 10.1097/00001756-200008210-00013. [DOI] [PubMed] [Google Scholar]
  4. Amalfitano A, et al. Production and characterization of improved adenovirus vectors with the E1, E2b, and E3 genes deleted. J Virol. 1998;72:926–33. doi: 10.1128/jvi.72.2.926-933.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Anderson DB, et al. Pseudotyping of glycoprotein D-deficient herpes simplex virus type 1 with vesicular stomatitis virus glycoprotein G enables mutant virus attachment and entry. J Virol. 2000;74:2481–7. doi: 10.1128/jvi.74.5.2481-2487.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Andreansky S, et al. Evaluation of genetically engineered herpes simplex viruses as oncolytic agents for human malignant brain tumors. Cancer Res. 1997;57:1502–9. [PubMed] [Google Scholar]
  7. Apolonia L, et al. Stable gene transfer to muscle using non-integrating lentiviral vectors. Mol Ther. 2007;15:1947–54. doi: 10.1038/sj.mt.6300281. [DOI] [PubMed] [Google Scholar]
  8. Armentano D, et al. Characterization of an adenovirus gene transfer vector containing an E4 deletion. Hum Gene Ther. 1995;6:1343–53. doi: 10.1089/hum.1995.6.10-1343. [DOI] [PubMed] [Google Scholar]
  9. Atchison RW, et al. Adenovirus-Associated Defective Virus Particles. Science. 1965;149:754–6. doi: 10.1126/science.149.3685.754. [DOI] [PubMed] [Google Scholar]
  10. Bayer M, et al. A large U3 deletion causes increased in vivo expression from a nonintegrating lentiviral vector. Mol Ther. 2008;16:1968–76. doi: 10.1038/mt.2008.199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Bearer EL, et al. Retrograde axonal transport of herpes simplex virus: evidence for a single mechanism and a role for tegument. Proc Natl Acad Sci U S A. 2000;97:8146–50. doi: 10.1073/pnas.97.14.8146. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Bell P, et al. Analysis of tumors arising in male B6C3F1 mice with and without AAV vector delivery to liver. Mol Ther. 2006;14:34–44. doi: 10.1016/j.ymthe.2006.03.008. [DOI] [PubMed] [Google Scholar]
  13. Bell P, et al. No evidence for tumorigenesis of AAV vectors in a large-scale study in mice. Mol Ther. 2005;12:299–306. doi: 10.1016/j.ymthe.2005.03.020. [DOI] [PubMed] [Google Scholar]
  14. Bergelson JM, et al. Isolation of a common receptor for Coxsackie B viruses and adenoviruses 2 and 5. Science. 1997;275:1320–3. doi: 10.1126/science.275.5304.1320. [DOI] [PubMed] [Google Scholar]
  15. Berk AJ. Adenoviridae: The Viruses and Their Replication. In: Knipe DM, Howley PM, editors. Fields virology. Wolters Kluwer Health/Lippincott Williams & Wilkins; Philadelphia: 2007. pp. 2355–2394. [Google Scholar]
  16. Berns KI, Parrish CR. Parvoviridae. In: Fields BN, et al., editors. Fields virology. Wolters Kluwer Health/Lippincott Williams & Wilkins; Philadelphia: 2007. pp. 2437–2477. [Google Scholar]
  17. Bessis N, et al. Immune responses to gene therapy vectors: influence on vector function and effector mechanisms. Gene Ther. 2004;11(Suppl 1):S10–7. doi: 10.1038/sj.gt.3302364. [DOI] [PubMed] [Google Scholar]
  18. Bett AJ, et al. An efficient and flexible system for construction of adenovirus vectors with insertions or deletions in early regions 1 and 3. Proc Natl Acad Sci U S A. 1994;91:8802–6. doi: 10.1073/pnas.91.19.8802. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Bischof D, Cornetta K. Flexibility in cell targeting by pseudotyping lentiviral vectors. Methods Mol Biol. 2010;614:53–68. doi: 10.1007/978-1-60761-533-0_3. [DOI] [PubMed] [Google Scholar]
  20. Boehmer PE, Lehman IR. Herpes simplex virus DNA replication. Annu Rev Biochem. 1997;66:347–84. doi: 10.1146/annurev.biochem.66.1.347. [DOI] [PubMed] [Google Scholar]
  21. Bokhoven M, et al. Insertional gene activation by lentiviral and gammaretroviral vectors. J Virol. 2009;83:283–94. doi: 10.1128/JVI.01865-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Briggs JA, et al. Structural organization of authentic, mature HIV-1 virions and cores. EMBO J. 2003;22:1707–15. doi: 10.1093/emboj/cdg143. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Broekman ML, et al. Adeno-associated virus vectors serotyped with AAV8 capsid are more efficient than AAV-1 or -2 serotypes for widespread gene delivery to the neonatal mouse brain. Neuroscience. 2006;138:501–10. doi: 10.1016/j.neuroscience.2005.11.057. [DOI] [PubMed] [Google Scholar]
  24. Brown PO, et al. Retroviral integration: structure of the initial covalent product and its precursor, and a role for the viral IN protein. Proc Natl Acad Sci U S A. 1989;86:2525–9. doi: 10.1073/pnas.86.8.2525. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Bukrinsky MI, et al. A nuclear localization signal within HIV-1 matrix protein that governs infection of non-dividing cells. Nature. 1993;365:666–9. doi: 10.1038/365666a0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Burger C, et al. 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. 2004;10:302–17. doi: 10.1016/j.ymthe.2004.05.024. [DOI] [PubMed] [Google Scholar]
  27. Candolfi M, et al. Effective high-capacity gutless adenoviral vectors mediate transgene expression in human glioma cells. Mol Ther. 2006;14:371–81. doi: 10.1016/j.ymthe.2006.05.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Candolfi M, et al. Optimization of adenoviral vector-mediated transgene expression in the canine brain in vivo, and in canine glioma cells in vitro. Neuro Oncol. 2007;9:245–58. doi: 10.1215/15228517-2007-012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Cara A, Reitz MS., Jr New insight on the role of extrachromosomal retroviral DNA. Leukemia. 1997;11:1395–9. doi: 10.1038/sj.leu.2400776. [DOI] [PubMed] [Google Scholar]
  30. Cassinotti P, et al. Organization of the adeno-associated virus (AAV) capsid gene: mapping of a minor spliced mRNA coding for virus capsid protein. Virology. 1988;167:176–84. [PubMed] [Google Scholar]
  31. Chao H, et al. Several log increase in therapeutic transgene delivery by distinct adeno-associated viral serotype vectors. Mol Ther. 2000;2:619–23. doi: 10.1006/mthe.2000.0219. [DOI] [PubMed] [Google Scholar]
  32. Chen J, et al. High efficiency of HIV-1 genomic RNA packaging and heterozygote formation revealed by single virion analysis. Proc Natl Acad Sci U S A. 2009a;106:13535–40. doi: 10.1073/pnas.0906822106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Chen ST, et al. Generation of packaging cell lines for pseudotyped retroviral vectors of the G protein of vesicular stomatitis virus by using a modified tetracycline inducible system. Proc Natl Acad Sci U S A. 1996;93:10057–62. doi: 10.1073/pnas.93.19.10057. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Chen YH, et al. Molecular signatures of disease brain endothelia provide new sites for CNS-directed enzyme therapy. Nat Med. 2009b;15:1215–8. doi: 10.1038/nm.2025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Cheung AK, et al. Integration of the adeno-associated virus genome into cellular DNA in latently infected human Detroit 6 cells. J Virol. 1980;33:739–48. doi: 10.1128/jvi.33.2.739-748.1980. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Chirmule N, et al. Immune responses to adenovirus and adeno-associated virus in humans. Gene Ther. 1999;6:1574–83. doi: 10.1038/sj.gt.3300994. [DOI] [PubMed] [Google Scholar]
  37. Cockrell AS, et al. A trans-lentiviral packaging cell line for high-titer conditional self-inactivating HIV-1 vectors. Mol Ther. 2006;14:276–84. doi: 10.1016/j.ymthe.2005.12.015. [DOI] [PubMed] [Google Scholar]
  38. Colin A, et al. Engineered lentiviral vector targeting astrocytes in vivo. Glia. 2009;57:667–79. doi: 10.1002/glia.20795. [DOI] [PubMed] [Google Scholar]
  39. Cornu TI, Cathomen T. Targeted genome modifications using integrase-deficient lentiviral vectors. Mol Ther. 2007;15:2107–13. doi: 10.1038/sj.mt.6300345. [DOI] [PubMed] [Google Scholar]
  40. Danos O, Mulligan RC. Safe and efficient generation of recombinant retroviruses with amphotropic and ecotropic host ranges. Proc Natl Acad Sci U S A. 1988;85:6460–4. doi: 10.1073/pnas.85.17.6460. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Davidson BL, et al. 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 U S A. 2000;97:3428–32. doi: 10.1073/pnas.050581197. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Davison AJ, et al. Genetic content and evolution of adenoviruses. J Gen Virol. 2003;84:2895–908. doi: 10.1099/vir.0.19497-0. [DOI] [PubMed] [Google Scholar]
  43. de Marco A, et al. Structural analysis of HIV-1 maturation using cryo-electron tomography. PLoS Pathog. 2010;6:e1001215. doi: 10.1371/journal.ppat.1001215. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Desmaris N, et al. Production and neurotropism of lentivirus vectors pseudotyped with lyssavirus envelope glycoproteins. Mol Ther. 2001;4:149–56. doi: 10.1006/mthe.2001.0431. [DOI] [PubMed] [Google Scholar]
  45. Dong B, et al. Characterization of genome integrity for oversized recombinant AAV vector. Mol Ther. 2010;18:87–92. doi: 10.1038/mt.2009.258. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Dong JY, et al. Quantitative analysis of the packaging capacity of recombinant adeno-associated virus. Hum Gene Ther. 1996;7:2101–12. doi: 10.1089/hum.1996.7.17-2101. [DOI] [PubMed] [Google Scholar]
  47. Duan D, et al. Circular intermediates of recombinant adeno-associated virus have defined structural characteristics responsible for long-term episomal persistence in muscle tissue. J Virol. 1998;72:8568–77. doi: 10.1128/jvi.72.11.8568-8577.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Duan D, et al. Expanding AAV packaging capacity with trans-splicing or overlapping vectors: a quantitative comparison. Mol Ther. 2001;4:383–91. doi: 10.1006/mthe.2001.0456. [DOI] [PubMed] [Google Scholar]
  49. Duan D, et al. A new dual-vector approach to enhance recombinant adeno-associated virus-mediated gene expression through intermolecular cis activation. Nat Med. 2000;6:595–8. doi: 10.1038/75080. [DOI] [PubMed] [Google Scholar]
  50. Dull T, et al. A third-generation lentivirus vector with a conditional packaging system. J Virol. 1998;72:8463–71. doi: 10.1128/jvi.72.11.8463-8471.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Duque S, et al. Intravenous administration of self-complementary AAV9 enables transgene delivery to adult motor neurons. Mol Ther. 2009;17:1187–96. doi: 10.1038/mt.2009.71. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Epstein AL. HSV-1-derived amplicon vectors: recent technological improvements and remaining difficulties--a review. Mem Inst Oswaldo Cruz. 2009;104:399–410. doi: 10.1590/s0074-02762009000300002. [DOI] [PubMed] [Google Scholar]
  53. Escors D, Breckpot K. Lentiviral vectors in gene therapy: their current status and future potential. Arch Immunol Ther Exp (Warsz) 2010;58:107–19. doi: 10.1007/s00005-010-0063-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Everts M, Curiel DT. Transductional targeting of adenoviral cancer gene therapy. Curr Gene Ther. 2004;4:337–46. doi: 10.2174/1566523043346372. [DOI] [PubMed] [Google Scholar]
  55. Ferrari FK, et al. Second-strand synthesis is a rate-limiting step for efficient transduction by recombinant adeno-associated virus vectors. J Virol. 1996;70:3227–34. doi: 10.1128/jvi.70.5.3227-3234.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Forster F, et al. Retrovirus envelope protein complex structure in situ studied by cryo-electron tomography. Proc Natl Acad Sci U S A. 2005;102:4729–34. doi: 10.1073/pnas.0409178102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Foust KD, Kaspar BK. Over the barrier and through the blood: to CNS delivery we go. Cell Cycle. 2009;8:4017–8. doi: 10.4161/cc.8.24.10245. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Foust KD, et al. Intravascular AAV9 preferentially targets neonatal neurons and adult astrocytes. Nat Biotechnol. 2009;27:59–65. doi: 10.1038/nbt.1515. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Fraefel C, et al. Helper virus-free transfer of herpes simplex virus type 1 plasmid vectors into neural cells. J Virol. 1996;70:7190–7. doi: 10.1128/jvi.70.10.7190-7197.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Freed EO. HIV-1 gag proteins: diverse functions in the virus life cycle. Virology. 1998;251:1–15. doi: 10.1006/viro.1998.9398. [DOI] [PubMed] [Google Scholar]
  61. Friedmann T, Roblin R. Gene therapy for human genetic disease? Science. 1972;175:949–55. doi: 10.1126/science.175.4025.949. [DOI] [PubMed] [Google Scholar]
  62. Garber DA, et al. Demonstration of circularization of herpes simplex virus DNA following infection using pulsed field gel electrophoresis. Virology. 1993;197:459–62. doi: 10.1006/viro.1993.1612. [DOI] [PubMed] [Google Scholar]
  63. Germano IM, et al. Adenovirus/herpes simplex-thymidine kinase/ganciclovir complex: preliminary results of a phase I trial in patients with recurrent malignant gliomas. J Neurooncol. 2003;65:279–89. doi: 10.1023/b:neon.0000003657.95085.56. [DOI] [PubMed] [Google Scholar]
  64. Ghosh A, et al. Efficient transgene reconstitution with hybrid dual AAV vectors carrying the minimized bridging sequences. Hum Gene Ther. 2011;22:77–83. doi: 10.1089/hum.2010.122. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Gijsbers R, et al. LEDGF hybrids efficiently retarget lentiviral integration into heterochromatin. Mol Ther. 2010;18:552–60. doi: 10.1038/mt.2010.36. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Glasgow JN, et al. Transductional targeting of adenovirus vectors for gene therapy. Cancer Gene Ther. 2006;13:830–44. doi: 10.1038/sj.cgt.7700928. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Goff SP. Retroviridae: The Retroviruses and Their Replication. In: Knipe DM, Howley PM, editors. Fields Virology. Wolters Kluwer Health/Lippincott Williams & Wilkins; Philadelphia: 2007. pp. 2000–2069. [Google Scholar]
  68. Gray SJ, et al. Directed evolution of a novel adeno-associated virus (AAV) vector that crosses the seizure-compromised blood-brain barrier (BBB) Mol Ther. 2010a;18:570–8. doi: 10.1038/mt.2009.292. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Gray SJ, et al. Preclinical differences of intravascular AAV9 delivery to neurons and glia: a comparative study of adult mice and nonhuman primates. Mol Ther. 2011;19:1058–69. doi: 10.1038/mt.2011.72. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Gray SJ, et al. Viral vectors and delivery strategies for CNS gene therapy. Therapeutic Delivery. 2010b;1:517–534. doi: 10.4155/tde.10.50. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Greenberg KP, et al. Targeted transgene expression in muller glia of normal and diseased retinas using lentiviral vectors. Invest Ophthalmol Vis Sci. 2007;48:1844–52. doi: 10.1167/iovs.05-1570. [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Grieger JC, et al. Production and characterization of adeno-associated viral vectors. Nat Protoc. 2006;1:1412–28. doi: 10.1038/nprot.2006.207. [DOI] [PubMed] [Google Scholar]
  73. Griffiths RA, et al. Herpesvirus saimiri-based gene delivery vectors. Curr Gene Ther. 2006;6:1–15. doi: 10.2174/156652306775515529. [DOI] [PubMed] [Google Scholar]
  74. Grunewald K, et al. Three-dimensional structure of herpes simplex virus from cryo-electron tomography. Science. 2003;302:1396–8. doi: 10.1126/science.1090284. [DOI] [PubMed] [Google Scholar]
  75. Hacein-Bey-Abina S, et al. LMO2-associated clonal T cell proliferation in two patients after gene therapy for SCID-X1. Science. 2003;302:415–9. doi: 10.1126/science.1088547. [DOI] [PubMed] [Google Scholar]
  76. Hearing P, et al. Identification of a repeated sequence element required for efficient encapsidation of the adenovirus type 5 chromosome. J Virol. 1987;61:2555–8. doi: 10.1128/jvi.61.8.2555-2558.1987. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Hewitt FC, et al. Reducing the risk of adeno-associated virus (AAV) vector mobilization with AAV type 5 vectors. J Virol. 2009;83:3919–29. doi: 10.1128/JVI.02466-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Hewitt FC, Samulski RJ. Creating a novel origin of replication through modulating DNA-protein interfaces. PLoS One. 2010;5:e8850. doi: 10.1371/journal.pone.0008850. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Hillgenberg M, et al. Chromosomal integration pattern of a helper-dependent minimal adenovirus vector with a selectable marker inserted into a 27.4-kilobase genomic stuffer. J Virol. 2001;75:9896–908. doi: 10.1128/JVI.75.20.9896-9908.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Hirsch ML, et al. Little vector, big gene transduction: fragmented genome reassembly of adeno-associated virus. Mol Ther. 2010;18:6–8. doi: 10.1038/mt.2009.280. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Hirsch ML, et al. AAV recombineering with single strand oligonucleotides. PLoS One. 2009;4:e7705. doi: 10.1371/journal.pone.0007705. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Im DS, Muzyczka N. The AAV origin binding protein Rep68 is an ATP-dependent site-specific endonuclease with DNA helicase activity. Cell. 1990;61:447–57. doi: 10.1016/0092-8674(90)90526-k. [DOI] [PubMed] [Google Scholar]
  83. Im DS, Muzyczka N. Partial purification of adeno-associated virus Rep78, Rep52, and Rep40 and their biochemical characterization. J Virol. 1992;66:1119–28. doi: 10.1128/jvi.66.2.1119-1128.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Immonen A, et al. AdvHSV-tk gene therapy with intravenous ganciclovir improves survival in human malignant glioma: a randomised, controlled study. Mol Ther. 2004;10:967–72. doi: 10.1016/j.ymthe.2004.08.002. [DOI] [PubMed] [Google Scholar]
  85. Jager L, Ehrhardt A. Persistence of high-capacity adenoviral vectors as replication-defective monomeric genomes in vitro and in murine liver. Hum Gene Ther. 2009;20:883–96. doi: 10.1089/hum.2009.020. [DOI] [PubMed] [Google Scholar]
  86. Jakobsson J, Lundberg C. Lentiviral vectors for use in the central nervous system. Mol Ther. 2006;13:484–93. doi: 10.1016/j.ymthe.2005.11.012. [DOI] [PubMed] [Google Scholar]
  87. Jakobsson J, et al. Efficient transduction of neurons using Ross River glycoprotein-pseudotyped lentiviral vectors. Gene Ther. 2006;13:966–73. doi: 10.1038/sj.gt.3302701. [DOI] [PubMed] [Google Scholar]
  88. Johnson MC. Mechanisms for Env glycoprotein acquisition by retroviruses. AIDS Res Hum Retroviruses. 2011;27:239–47. doi: 10.1089/aid.2010.0350. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Kafri T, et al. A packaging cell line for lentivirus vectors. J Virol. 1999;73:576–84. doi: 10.1128/jvi.73.1.576-584.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Kaludov N, et al. 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;75:6884–93. doi: 10.1128/JVI.75.15.6884-6893.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Kantor B, et al. Notable Reduction in Illegitimate Integration Mediated by a PPT-deleted, Nonintegrating Lentiviral Vector. Mol Ther. 2011;19:547–56. doi: 10.1038/mt.2010.277. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Kaplitt MG, et al. Long-term gene expression and phenotypic correction using adeno-associated virus vectors in the mammalian brain. Nat Genet. 1994;8:148–54. doi: 10.1038/ng1094-148. [DOI] [PubMed] [Google Scholar]
  93. Kay MA. AAV vectors and tumorigenicity. Nat Biotechnol. 2007;25:1111–3. doi: 10.1038/nbt1007-1111. [DOI] [PubMed] [Google Scholar]
  94. Klages N, et al. A stable system for the high-titer production of multiply attenuated lentiviral vectors. Mol Ther. 2000;2:170–6. doi: 10.1006/mthe.2000.0103. [DOI] [PubMed] [Google Scholar]
  95. Klein RL, et al. Long-term actions of vector-derived nerve growth factor or brain-derived neurotrophic factor on choline acetyltransferase and Trk receptor levels in the adult rat basal forebrain. Neuroscience. 1999;90:815–21. doi: 10.1016/s0306-4522(98)00537-5. [DOI] [PubMed] [Google Scholar]
  96. Koerber JT, et al. Molecular evolution of adeno-associated virus for enhanced glial gene delivery. Mol Ther. 2009;17:2088–95. doi: 10.1038/mt.2009.184. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Kreppel F, Kochanek S. Long-term transgene expression in proliferating cells mediated by episomally maintained high-capacity adenovirus vectors. J Virol. 2004;78:9–22. doi: 10.1128/JVI.78.1.9-22.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Krisky DM, et al. Development of herpes simplex virus replication-defective multigene vectors for combination gene therapy applications. Gene Ther. 1998a;5:1517–30. doi: 10.1038/sj.gt.3300755. [DOI] [PubMed] [Google Scholar]
  99. Krisky DM, et al. Deletion of multiple immediate-early genes from herpes simplex virus reduces cytotoxicity and permits long-term gene expression in neurons. Gene Ther. 1998b;5:1593–603. doi: 10.1038/sj.gt.3300766. [DOI] [PubMed] [Google Scholar]
  100. Kumar M, et al. Systematic determination of the packaging limit of lentiviral vectors. Hum Gene Ther. 2001;12:1893–905. doi: 10.1089/104303401753153947. [DOI] [PubMed] [Google Scholar]
  101. Lai Y, et al. Evidence for the failure of adeno-associated virus serotype 5 to package a viral genome > or = 8.2 kb. Mol Ther. 2010;18:75–9. doi: 10.1038/mt.2009.256. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Lawlor PA, et al. Efficient gene delivery and selective transduction of glial cells in the mammalian brain by AAV serotypes isolated from nonhuman primates. Mol Ther. 2009;17:1692–702. doi: 10.1038/mt.2009.170. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Leis J, et al. Standardized and simplified nomenclature for proteins common to all retroviruses. J Virol. 1988;62:1808–9. doi: 10.1128/jvi.62.5.1808-1809.1988. [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Lewis PF, Emerman M. Passage through mitosis is required for oncoretroviruses but not for the human immunodeficiency virus. J Virol. 1994;68:510–6. doi: 10.1128/jvi.68.1.510-516.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Lilley CE, et al. Multiple immediate-early gene-deficient herpes simplex virus vectors allowing efficient gene delivery to neurons in culture and widespread gene delivery to the central nervous system in vivo. J Virol. 2001;75:4343–56. doi: 10.1128/JVI.75.9.4343-4356.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Liu J, et al. 3D visualization of HIV virions by cryoelectron tomography. Methods Enzymol. 2010;483:267–90. doi: 10.1016/S0076-6879(10)83014-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  107. Lock M, et al. Rapid, simple, and versatile manufacturing of recombinant adeno-associated viral vectors at scale. Hum Gene Ther. 2010;21:1259–71. doi: 10.1089/hum.2010.055. [DOI] [PMC free article] [PubMed] [Google Scholar]
  108. Lou E. Oncolytic herpes viruses as a potential mechanism for cancer therapy. Acta Oncol. 2003;42:660–71. doi: 10.1080/0284186031000518. [DOI] [PubMed] [Google Scholar]
  109. Lozier JN, et al. Toxicity of a first-generation adenoviral vector in rhesus macaques. Hum Gene Ther. 2002;13:113–24. doi: 10.1089/10430340152712665. [DOI] [PubMed] [Google Scholar]
  110. Lundberg C, et al. Applications of lentiviral vectors for biology and gene therapy of neurological disorders. Curr Gene Ther. 2008;8:461–73. doi: 10.2174/156652308786847996. [DOI] [PubMed] [Google Scholar]
  111. Lusby E, et al. Nucleotide sequence of the inverted terminal repetition in adeno-associated virus DNA. J Virol. 1980;34:402–9. doi: 10.1128/jvi.34.2.402-409.1980. [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Maier P, et al. Retroviral vectors for gene therapy. Future Microbiol. 2010;5:1507–23. doi: 10.2217/fmb.10.100. [DOI] [PubMed] [Google Scholar]
  113. Marconi P, et al. HSV as a vector in vaccine development and gene therapy. Adv Exp Med Biol. 2009;655:118–44. doi: 10.1007/978-1-4419-1132-2_10. [DOI] [PubMed] [Google Scholar]
  114. Marconi P, et al. Replication-defective herpes simplex virus vectors for gene transfer in vivo. Proc Natl Acad Sci U S A. 1996;93:11319–20. doi: 10.1073/pnas.93.21.11319. [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Markert JM, et al. Phase Ib trial of mutant herpes simplex virus G207 inoculated pre-and post-tumor resection for recurrent GBM. Mol Ther. 2009;17:199–207. doi: 10.1038/mt.2008.228. [DOI] [PMC free article] [PubMed] [Google Scholar]
  116. Markert JM, et al. Conditionally replicating herpes simplex virus mutant, G207 for the treatment of malignant glioma: results of a phase I trial. Gene Ther. 2000;7:867–74. doi: 10.1038/sj.gt.3301205. [DOI] [PubMed] [Google Scholar]
  117. Markowitz D, et al. A safe packaging line for gene transfer: separating viral genes on two different plasmids. J Virol. 1988;62:1120–4. doi: 10.1128/jvi.62.4.1120-1124.1988. [DOI] [PMC free article] [PubMed] [Google Scholar]
  118. McCarty DM, et al. Self-complementary recombinant adeno-associated virus (scAAV) vectors promote efficient transduction independently of DNA synthesis. Gene Ther. 2001;8:1248–54. doi: 10.1038/sj.gt.3301514. [DOI] [PubMed] [Google Scholar]
  119. McCarty DM, et al. Integration of adeno-associated virus (AAV) and recombinant AAV vectors. Annu Rev Genet. 2004;38:819–45. doi: 10.1146/annurev.genet.37.110801.143717. [DOI] [PubMed] [Google Scholar]
  120. McCown TJ. Adeno-Associated Virus (AAV) Vectors in the CNS. Curr Gene Ther. 2011;11:181–8. doi: 10.2174/156652311795684759. [DOI] [PubMed] [Google Scholar]
  121. Michelfelder S, Trepel M. Adeno-associated viral vectors and their redirection to cell-type specific receptors. Adv Genet. 2009;67:29–60. doi: 10.1016/S0065-2660(09)67002-4. [DOI] [PubMed] [Google Scholar]
  122. Miller AD. Cell-surface receptors for retroviruses and implications for gene transfer. Proc Natl Acad Sci U S A. 1996;93:11407–13. doi: 10.1073/pnas.93.21.11407. [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Mineta T, et al. Attenuated multi-mutated herpes simplex virus-1 for the treatment of malignant gliomas. Nat Med. 1995;1:938–43. doi: 10.1038/nm0995-938. [DOI] [PubMed] [Google Scholar]
  124. Miyoshi H, et al. Development of a self-inactivating lentivirus vector. J Virol. 1998;72:8150–7. doi: 10.1128/jvi.72.10.8150-8157.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  125. Morral N, et al. Administration of helper-dependent adenoviral vectors and sequential delivery of different vector serotype for long-term liver-directed gene transfer in baboons. Proc Natl Acad Sci U S A. 1999;96:12816–21. doi: 10.1073/pnas.96.22.12816. [DOI] [PMC free article] [PubMed] [Google Scholar]
  126. Morsy MA, et al. An adenoviral vector deleted for all viral coding sequences results in enhanced safety and extended expression of a leptin transgene. Proc Natl Acad Sci U S A. 1998;95:7866–71. doi: 10.1073/pnas.95.14.7866. [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Muzyczka N, Warrington KH., Jr Custom adeno-associated virus capsids: the next generation of recombinant vectors with novel tropism. Hum Gene Ther. 2005;16:408–16. doi: 10.1089/hum.2005.16.408. [DOI] [PubMed] [Google Scholar]
  128. Nakai H, et al. Increasing the size of rAAV-mediated expression cassettes in vivo by intermolecular joining of two complementary vectors. Nat Biotechnol. 2000;18:527–32. doi: 10.1038/75390. [DOI] [PubMed] [Google Scholar]
  129. Naldini L, et al. In vivo gene delivery and stable transduction of nondividing cells by a lentiviral vector. Science. 1996;272:263–7. doi: 10.1126/science.272.5259.263. [DOI] [PubMed] [Google Scholar]
  130. Nash K, et al. Complete in vitro reconstitution of adeno-associated virus DNA replication requires the minichromosome maintenance complex proteins. J Virol. 2008;82:1458–64. doi: 10.1128/JVI.01968-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  131. Niemeyer GP, et al. Long-term correction of inhibitor-prone hemophilia B dogs treated with liver-directed AAV2-mediated factor IX gene therapy. Blood. 2009;113:797–806. doi: 10.1182/blood-2008-10-181479. [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. NIH Guidelines for Research Involving Recombinant DNA Molecules (NIH Guidelines) http://oba.od.nih.gov/oba/rac/Guidelines/NIH_Guidelines.htm.
  133. Oxford Biomedica. Oxford BioMedica Announces Interim Data from Highest (5×) Dose Cohort in ProSavin® Phase I/II Study in Parkinson’s Disease - 04/08/2011. 2011 http://www.oxfordbiomedica.co.uk/page.asp?pageid=59&newsid=297.
  134. Pakzaban P, et al. Effect of exogenous nerve growth factor on neurotoxicity of and neuronal gene delivery by a herpes simplex amplicon vector in the rat brain. Hum Gene Ther. 1994;5:987–95. doi: 10.1089/hum.1994.5.8-987. [DOI] [PubMed] [Google Scholar]
  135. Parks RJ. Improvements in adenoviral vector technology: overcoming barriers for gene therapy. Clin Genet. 2000;58:1–11. doi: 10.1034/j.1399-0004.2000.580101.x. [DOI] [PubMed] [Google Scholar]
  136. Parks RJ, et al. A helper-dependent adenovirus vector system: removal of helper virus by Cre-mediated excision of the viral packaging signal. Proc Natl Acad Sci U S A. 1996;93:13565–70. doi: 10.1073/pnas.93.24.13565. [DOI] [PMC free article] [PubMed] [Google Scholar]
  137. Pellett PE, Roizman B. The Family Herpesviridae: A Brief Introduction. In: Knipe DM, Howley PM, editors. Fields Virology. Wolters Kluwer Health/Lippincott Williams & Wilkins; Philadelphia: 2007. pp. 2479–2499. [Google Scholar]
  138. Philippe S, et al. Lentiviral vectors with a defective integrase allow efficient and sustained transgene expression in vitro and in vivo. Proc Natl Acad Sci U S A. 2006;103:17684–9. doi: 10.1073/pnas.0606197103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  139. Philpott NJ, Thrasher AJ. Use of nonintegrating lentiviral vectors for gene therapy. Hum Gene Ther. 2007;18:483–9. doi: 10.1089/hum.2007.013. [DOI] [PubMed] [Google Scholar]
  140. Pike-Overzet K, et al. New insights and unresolved issues regarding insertional mutagenesis in X-linked SCID gene therapy. Mol Ther. 2007;15:1910–6. doi: 10.1038/sj.mt.6300297. [DOI] [PubMed] [Google Scholar]
  141. Rabinowitz JE, et al. Cross-packaging of a single adeno-associated virus (AAV) type 2 vector genome into multiple AAV serotypes enables transduction with broad specificity. J Virol. 2002;76:791–801. doi: 10.1128/JVI.76.2.791-801.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Rahim AA, et al. Efficient gene delivery to the adult and fetal CNS using pseudotyped non-integrating lentiviral vectors. Gene Ther. 2009;16:509–20. doi: 10.1038/gt.2008.186. [DOI] [PubMed] [Google Scholar]
  143. Rauschhuber C, et al. New insights into stability of recombinant adenovirus vector genomes in mammalian cells. Eur J Cell Biol. 2011 doi: 10.1016/j.ejcb.2011.01.006. [DOI] [PubMed] [Google Scholar]
  144. Reddy VS, et al. Crystal structure of human adenovirus at 3.5 A resolution. Science. 2010;329:1071–5. doi: 10.1126/science.1187292. [DOI] [PMC free article] [PubMed] [Google Scholar]
  145. Rivera VM, et al. Long-term pharmacologically regulated expression of erythropoietin in primates following AAV-mediated gene transfer. Blood. 2005;105:1424–30. doi: 10.1182/blood-2004-06-2501. [DOI] [PubMed] [Google Scholar]
  146. Robertson ES, et al. Epstein-Barr virus vectors for gene delivery to B lymphocytes. Proc Natl Acad Sci U S A. 1996;93:11334–40. doi: 10.1073/pnas.93.21.11334. [DOI] [PMC free article] [PubMed] [Google Scholar]
  147. Roizman B, et al. Herpes Simplex Viruses. In: Knipe DM, Howley PM, editors. Fields Virology. Wolters Kluwer Health/Lippincott Williams & Wilkins; Philadelphia: 2007. pp. 2502–2601. [Google Scholar]
  148. Rose JA, et al. Structural proteins of adenovirus-associated viruses. J Virol. 1971;8:766–70. doi: 10.1128/jvi.8.5.766-770.1971. [DOI] [PMC free article] [PubMed] [Google Scholar]
  149. Saeki Y, et al. Improved HSV-1 amplicon packaging system using ICP27-deleted, oversized HSV-1 BAC DNA. Methods Mol Med. 2003;76:51–60. doi: 10.1385/1-59259-304-6:51. [DOI] [PubMed] [Google Scholar]
  150. Samaniego LA, et al. Persistence and expression of the herpes simplex virus genome in the absence of immediate-early proteins. J Virol. 1998;72:3307–20. doi: 10.1128/jvi.72.4.3307-3320.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  151. Samulski RJ, et al. Helper-free stocks of recombinant adeno-associated viruses: normal integration does not require viral gene expression. J Virol. 1989;63:3822–8. doi: 10.1128/jvi.63.9.3822-3828.1989. [DOI] [PMC free article] [PubMed] [Google Scholar]
  152. Sarkis C, et al. Non-integrating lentiviral vectors. Curr Gene Ther. 2008;8:430–7. doi: 10.2174/156652308786848012. [DOI] [PubMed] [Google Scholar]
  153. Scarlata S, Carter C. Role of HIV-1 Gag domains in viral assembly. Biochim Biophys Acta. 2003;1614:62–72. doi: 10.1016/s0005-2736(03)00163-9. [DOI] [PubMed] [Google Scholar]
  154. Schnepp BC, et al. Characterization of adeno-associated virus genomes isolated from human tissues. J Virol. 2005;79:14793–803. doi: 10.1128/JVI.79.23.14793-14803.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  155. Shen S, et al. Terminal N-Linked Galactose Is the Primary Receptor for Adeno-associated Virus 9. J Biol Chem. 2011;286:13532–40. doi: 10.1074/jbc.M110.210922. [DOI] [PMC free article] [PubMed] [Google Scholar]
  156. Shen Y, Post L. Viral Vectors and Their Applications. In: Fields BN, et al., editors. Fields virology. Wolters Kluwer Health/Lippincott Williams & Wilkins; Philadelphia: 2007. pp. 539–564. [Google Scholar]
  157. Shieh MT, et al. Cell surface receptors for herpes simplex virus are heparan sulfate proteoglycans. J Cell Biol. 1992;116:1273–81. doi: 10.1083/jcb.116.5.1273. [DOI] [PMC free article] [PubMed] [Google Scholar]
  158. Snyder RO, et al. In vitro resolution of covalently joined AAV chromosome ends. Cell. 1990;60:105–13. doi: 10.1016/0092-8674(90)90720-y. [DOI] [PubMed] [Google Scholar]
  159. Sommerfelt MA. Retrovirus receptors. J Gen Virol. 1999;80(Pt 12):3049–64. doi: 10.1099/0022-1317-80-12-3049. [DOI] [PubMed] [Google Scholar]
  160. Spaete RR, Frenkel N. The herpes simplex virus amplicon: a new eucaryotic defective-virus cloning-amplifying vector. Cell. 1982;30:295–304. doi: 10.1016/0092-8674(82)90035-6. [DOI] [PubMed] [Google Scholar]
  161. Stephen SL, et al. Chromosomal integration of adenoviral vector DNA in vivo. J Virol. 2010;84:9987–94. doi: 10.1128/JVI.00751-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  162. Stephen SL, et al. Homologous and heterologous recombination between adenovirus vector DNA and chromosomal DNA. J Gene Med. 2008;10:1176–89. doi: 10.1002/jgm.1246. [DOI] [PubMed] [Google Scholar]
  163. Stieger K, et al. Detection of intact rAAV particles up to 6 years after successful gene transfer in the retina of dogs and primates. Mol Ther. 2009;17:516–23. doi: 10.1038/mt.2008.283. [DOI] [PMC free article] [PubMed] [Google Scholar]
  164. Strang BL, Stow ND. Circularization of the herpes simplex virus type 1 genome upon lytic infection. J Virol. 2005;79:12487–94. doi: 10.1128/JVI.79.19.12487-12494.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  165. Straus SE, et al. Concatemers of alternating plus and minus strands are intermediates in adenovirus-associated virus DNA synthesis. Proc Natl Acad Sci U S A. 1976;73:742–6. doi: 10.1073/pnas.73.3.742. [DOI] [PMC free article] [PubMed] [Google Scholar]
  166. Summerford C, Samulski RJ. Membrane-associated heparan sulfate proteoglycan is a receptor for adeno-associated virus type 2 virions. J Virol. 1998;72:1438–45. doi: 10.1128/jvi.72.2.1438-1445.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  167. Sun L, et al. Overcoming adeno-associated virus vector size limitation through viral DNA heterodimerization. Nat Med. 2000;6:599–602. doi: 10.1038/75087. [DOI] [PubMed] [Google Scholar]
  168. Sun M, et al. Comparison of the capability of GDNF, BDNF, or both, to protect nigrostriatal neurons in a rat model of Parkinson’s disease. Brain Res. 2005;1052:119–29. doi: 10.1016/j.brainres.2005.05.072. [DOI] [PMC free article] [PubMed] [Google Scholar]
  169. Sun M, et al. Correction of a rat model of Parkinson’s disease by coexpression of tyrosine hydroxylase and aromatic amino acid decarboxylase from a helper virus-free herpes simplex virus type 1 vector. Hum Gene Ther. 2003;14:415–24. doi: 10.1089/104303403321467180. [DOI] [PMC free article] [PubMed] [Google Scholar]
  170. Suzuki T, et al. Development of a recombinant adenovirus vector production system free of replication-competent adenovirus by utilizing a packaging size limit of the viral genome. Virus Res. 2011 doi: 10.1016/j.virusres.2011.03.026. [DOI] [PubMed] [Google Scholar]
  171. Tattersall P, Ward DC. Rolling hairpin model for replication of parvovirus and linear chromosomal DNA. Nature. 1976;263:106–9. doi: 10.1038/263106a0. [DOI] [PubMed] [Google Scholar]
  172. Taymans JM, et al. Comparative analysis of adeno-associated viral vector serotypes 1, 2, 5, 7, and 8 in mouse brain. Hum Gene Ther. 2007;18:195–206. doi: 10.1089/hum.2006.178. [DOI] [PubMed] [Google Scholar]
  173. Telesnitsky A, Goff SP. Reverse Transcriptase and the Generation of Retroviral DNA. In: CJM, et al., editors. Retroviruses. Cold Spring Harbor Laboratory Press; Cold Spring Harbor (NY): 1997. [PubMed] [Google Scholar]
  174. Thomas CE, et al. Acute direct adenoviral vector cytotoxicity and chronic, but not acute, inflammatory responses correlate with decreased vector-mediated transgene expression in the brain. Mol Ther. 2001;3:36–46. doi: 10.1006/mthe.2000.0224. [DOI] [PubMed] [Google Scholar]
  175. Thomas CE, et al. Adenovirus binding to the coxsackievirus and adenovirus receptor or integrins is not required to elicit brain inflammation but is necessary to transduce specific neural cell types. J Virol. 2002;76:3452–60. doi: 10.1128/JVI.76.7.3452-3460.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  176. Tomko RP, et al. HCAR and MCAR: the human and mouse cellular receptors for subgroup C adenoviruses and group B coxsackieviruses. Proc Natl Acad Sci U S A. 1997;94:3352–6. doi: 10.1073/pnas.94.7.3352. [DOI] [PMC free article] [PubMed] [Google Scholar]
  177. Tuszynski MH, et al. A phase 1 clinical trial of nerve growth factor gene therapy for Alzheimer disease. Nat Med. 2005;11:551–5. doi: 10.1038/nm1239. [DOI] [PubMed] [Google Scholar]
  178. Van der Perren A, et al. Efficient and stable transduction of dopaminergic neurons in rat substantia nigra by rAAV 2/1, 2/2, 2/5, 2/6.2, 2/7, 2/8 and 2/9. Gene Ther. 2011;18:517–27. doi: 10.1038/gt.2010.179. [DOI] [PubMed] [Google Scholar]
  179. Van Vliet KM, et al. The role of the adeno-associated virus capsid in gene transfer. Methods Mol Biol. 2008;437:51–91. doi: 10.1007/978-1-59745-210-6_2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  180. Vandenberghe LH, et al. Heparin binding directs activation of T cells against adeno-associated virus serotype 2 capsid. Nat Med. 2006;12:967–71. doi: 10.1038/nm1445. [DOI] [PubMed] [Google Scholar]
  181. Vandenberghe LH, et al. Efficient serotype-dependent release of functional vector into the culture medium during adeno-associated virus manufacturing. Hum Gene Ther. 2010;21:1251–7. doi: 10.1089/hum.2010.107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  182. Volpers C, Kochanek S. Adenoviral vectors for gene transfer and therapy. J Gene Med. 2004;6(Suppl 1):S164–71. doi: 10.1002/jgm.496. [DOI] [PubMed] [Google Scholar]
  183. Wadsworth S, et al. Anatomy of herpes simplex virus DNA. II. Size, composition, and arrangement of inverted terminal repetitions. J Virol. 1975;15:1487–97. doi: 10.1128/jvi.15.6.1487-1497.1975. [DOI] [PMC free article] [PubMed] [Google Scholar]
  184. Wang C, et al. Recombinant AAV serotype 1 transduction efficiency and tropism in the murine brain. Gene Ther. 2003;10:1528–34. doi: 10.1038/sj.gt.3302011. [DOI] [PubMed] [Google Scholar]
  185. Weiss RA, Tailor CS. Retrovirus receptors. Cell. 1995;82:531–3. doi: 10.1016/0092-8674(95)90024-1. [DOI] [PubMed] [Google Scholar]
  186. Wiley Gene Therapy Clinical Trials Worldwide. 2011 http://www.wiley.com//legacy/wileychi/genmed/clinical/
  187. Wolfe JH, et al. Herpesvirus vector gene transfer and expression of beta-glucuronidase in the central nervous system of MPS VII mice. Nat Genet. 1992;1:379–84. doi: 10.1038/ng0892-379. [DOI] [PubMed] [Google Scholar]
  188. Wong LF, et al. Lentivirus-mediated gene transfer to the central nervous system: therapeutic and research applications. Hum Gene Ther. 2006;17:1–9. doi: 10.1089/hum.2006.17.1. [DOI] [PubMed] [Google Scholar]
  189. Wu N, et al. Prolonged gene expression and cell survival after infection by a herpes simplex virus mutant defective in the immediate-early genes encoding ICP4, ICP27, and ICP22. J Virol. 1996;70:6358–69. doi: 10.1128/jvi.70.9.6358-6369.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  190. Wu Z, et al. Adeno-associated virus serotypes: vector toolkit for human gene therapy. Mol Ther. 2006a;14:316–27. doi: 10.1016/j.ymthe.2006.05.009. [DOI] [PubMed] [Google Scholar]
  191. Wu Z, et al. Alpha2,3 and alpha2,6 N-linked sialic acids facilitate efficient binding and transduction by adeno-associated virus types 1 and 6. J Virol. 2006b;80:9093–103. doi: 10.1128/JVI.00895-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  192. Wu Z, et al. Effect of genome size on AAV vector packaging. Mol Ther. 2010;18:80–6. doi: 10.1038/mt.2009.255. [DOI] [PMC free article] [PubMed] [Google Scholar]
  193. Xiao W, et al. Gene therapy vectors based on adeno-associated virus type 1. J Virol. 1999;73:3994–4003. doi: 10.1128/jvi.73.5.3994-4003.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  194. Xie Q, et al. The atomic structure of adeno-associated virus (AAV-2), a vector for human gene therapy. Proc Natl Acad Sci U S A. 2002;99:10405–10. doi: 10.1073/pnas.162250899. [DOI] [PMC free article] [PubMed] [Google Scholar]
  195. Yamada M, et al. Herpes simplex virus vector-mediated expression of Bcl-2 prevents 6-hydroxydopamine-induced degeneration of neurons in the substantia nigra in vivo. Proc Natl Acad Sci U S A. 1999;96:4078–83. doi: 10.1073/pnas.96.7.4078. [DOI] [PMC free article] [PubMed] [Google Scholar]
  196. Yan Z, et al. Trans-splicing vectors expand the utility of adeno-associated virus for gene therapy. Proc Natl Acad Sci U S A. 2000;97:6716–21. doi: 10.1073/pnas.97.12.6716. [DOI] [PMC free article] [PubMed] [Google Scholar]
  197. Yanez-Munoz RJ, et al. Effective gene therapy with nonintegrating lentiviral vectors. Nat Med. 2006;12:348–53. doi: 10.1038/nm1365. [DOI] [PubMed] [Google Scholar]
  198. Zhang GR, et al. A tyrosine hydroxylase-neurofilament chimeric promoter enhances long-term expression in rat forebrain neurons from helper virus-free HSV-1 vectors. Brain Res Mol Brain Res. 2000;84:17–31. doi: 10.1016/s0169-328x(00)00197-2. [DOI] [PubMed] [Google Scholar]
  199. Zufferey R, et al. Self-inactivating lentivirus vector for safe and efficient in vivo gene delivery. J Virol. 1998;72:9873–80. doi: 10.1128/jvi.72.12.9873-9880.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  200. Zufferey R, et al. Multiply attenuated lentiviral vector achieves efficient gene delivery in vivo. Nat Biotechnol. 1997;15:871–5. doi: 10.1038/nbt0997-871. [DOI] [PubMed] [Google Scholar]

RESOURCES