Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2009 Oct 22.
Published in final edited form as: ILAR J. 2009;50(2):187–198. doi: 10.1093/ilar.50.2.187

Gene Therapy in Large Animal Models of Muscular Dystrophy

Zejing Wang 1, Jeffrey S Chamberlain 2, Stephen J Tapscott 3, Rainer Storb 4
PMCID: PMC2765825  NIHMSID: NIHMS150972  PMID: 19293461

Abstract

The muscular dystrophies are a group of genetically and phenotypically heterogeneously inherited diseases characterized by progressive muscle wasting, which can lead to premature death in severe forms such as Duchenne muscular dystrophy (DMD). In many cases they are caused by the absence of proteins that are critical components of the dystrophin-glycoprotein complex, which links the cytoskeleton and the basal lamina. There is no effective treatment for these disorders at present, but several novel strategies for replacing or repairing the defective gene are in development, with early encouraging results from animal models. We review these strategies, which include the use of stem cells of different tissue origins, gene replacement therapies mediated by various viral vectors, and transcript repair treatments using exon skipping strategies. We comment on their advantages and on limitations that must be overcome before successful application to human patients. Our focus is on studies in a clinically relevant large canine model of DMD. Recent advances in the field suggest that effective therapies for muscular dystrophies are on the horizon. Because of the complex nature of these diseases, it may be necessary to combine multiple approaches to achieve a successful treatment.

Keywords: dog model, Duchenne muscular dystrophy (DMD), dystrophin, gene therapy, muscular dystrophies, stem cell, viral vector, antisense oligonucleotide, immunosuppression

Introduction

Muscular Dystrophies

Muscular dystrophies are a group of heterogeneous diseases that primarily affect striated muscles throughout the body. Many of these myopathies are caused by mutations in genes that encode for structural proteins that link the cytoskeleton of muscle fibers to the extracellular matrix. The absence of functional proteins results in destabilization of the muscle membrane, increased muscle fragility and degeneration, and progressive muscle wasting, all of which compromise patients’ mobility and, in the severe disease forms (Emery 2002) such as Duchenne muscular dystrophy (DMD1), lead to death.

DMD is a lethal X-linked recessive disease that affects 1 of 3,500 boys worldwide. It is caused by loss of the protein dystrophin, a critical component of the dystrophin-glycoprotein complex (DGC1) at the sarcolemma (Hoffman et al. 1987; Muntoni et al. 2003; Tyler 2003). The absence of dystrophin prevents assembly of the DGC, resulting in a functionally impaired sarcolemma; membranes then become highly susceptible to mechanical contraction-induced damage, which leads to cycles of myofiber necrosis and regeneration, and hence progressive loss of muscle mass. As muscle tissue is lost, it is gradually replaced by connective tissue and adipose cells (Foidart et al. 1981).

The clinical course of DMD is severe and progressive, although the disease phenotype and progression vary and may change over time. Affected individuals can be diagnosed at birth on the basis of elevated serum levels of muscle enzymes. They exhibit muscle weakness by age 5, lose independent ambulation, and succumb to respiratory failure or cardiomyopathy in their late teens or twenties (Muntoni et al. 2003; Tyler 2003). The disease differentially affects adjacent muscles and may even completely spare some muscles, such as the extraocular muscle (Khurana et al. 1995; Porter 1998). Increasing data suggest that secondary responses, such as inflammatory processes, may play major roles in promoting the pathology of dystrophin-deficient muscle through upregulation of major histocompatibility complex (MHC) molecules, various chemokines, and molecules necessary for the costimulation and eventual activation of T cells (Porter et al. 2003; Tidball and Wehling-Henricks 2005; Wiendl et al. 2005). Loss of calcium homeostasis may also be a cause of muscular dystrophy, but whether it is the primary cause of muscle fiber degradation or a secondary phenomenon resulting from fiber microlesions remains to be determined (Muntoni et al. 2003; Spencer and Mellgren 2002).

Despite the well-understood pathogenesis of DMD, the development of curative therapies targeting its primary causes remains a major challenge. However, several novel strategies have been the subject of preclinical studies and raise hope for the discovery of treatments for human patients. Here, we review results of cell- and viral vector–based gene replacement and repair as well as antisense oligonucleotide-mediated gene correction in a canine model of DMD.

Dystrophin and Dystrophin Deficiency

Dystrophin, the largest known gene, is located on the X chromosome, with 79 exons encoding a full-length 14,000 bp mRNA distributed over more than 2 million bases of genomic sequences. Seven promoters linked to five different first exons give rise to various isoforms in a tissue-specific manner (Hoffman et al. 1987, 1988; Koenig et al. 1988; Muntoni et al. 2003). The full-length dystrophin protein is a large rod-shaped protein with a molecular weight of 427 kDa. It has four functional domains: the amino-terminus contains an actin-binding domain for anchoring dystrophin to the cytoskeleton; the central rod domain contains 24 spectrin-like repeats that constitute a flexible and elastic region with actin-binding properties; and the cysteine-rich and carboxyl-terminal domains interact with the DGC members (Abmayr and Chamberlain 2006; Hoffman et al. 1988; Koenig et al. 1988; Muntoni et al. 2003; Rando 2001). While the full-length dystrophin is normally expressed in striated muscle, smooth muscle, and neurons, multiple smaller dystrophin isoforms are exclusively or predominantly expressed in various nonmuscle tissues (e.g., the retina, glia, liver, and kidney) through the use of internal promoters and alternative splicing events (Byers et al. 1993; Lidov et al. 1995; Muntoni et al. 2003). Thus, although dystrophin deficiency is primarily manifested in muscle tissue, it can also lead to cognitive impairment (Muntoni et al. 2003). To date, there is no evidence that it leads to pathological abnormalities in other tissues.

Various mutations in the dystrophin gene can cause dystrophin deficiency, which presents as DMD or the milder Becker muscular dystrophy (BMD1). Intragenic deletions, the most common mutations, occur in 60–65% of DMD and BMD patients, duplication in 5–15% of the cases, and nonsense point mutations and other small mutations account for the remaining 20–35% (Muntoni et al. 2003). Disease severity is not simply related to the extent of a deletion or duplication but rather depends on whether it disrupts the normal open reading frame and allows expression of the dystroglycan-binding domain. Generally, in-frame mutations result in truncated yet partly functional dystrophin proteins and are associated with BMD, while frame shift mutations, which result in unstable RNA and complete absence of the dystrophin protein, are associated with DMD (Kerr et al. 2001; Muntoni et al. 2003). Information from studies of genotype-phenotype relationships in humans with partial deletion mutations and from mdx transgenic mice have shown that deletions in the N-terminal or the dystroglycan-binding domains of the dystrophin cause more severe clinical phenotypes (Harper et al. 2002b; Koenig et al. 1989; Muntoni et al. 2003).

Dystrophin interacts with integral and peripheral membrane proteins including dystroglycan, syntrophin, sarcoglycan, sarcospan, and dystrobrevin, and which collectively constitute the dystrophin-glycoprotein complex. At the sarcolemma of striated and smooth muscles, the DGC spans the plasma membrane and provides a strong mechanical link connecting the intracellular γ-actin cytoskeleton to the extracellular matrix (Ervasti 2006). The absence of dystrophin prevents assembly of the DGC and reduces the levels of all DGC components (Lapidos et al. 2004; Muntoni et al. 2003). The resulting sarcolemma is mechanically fragile due to the inability to laterally transmit forces from within myofibers to the extracellular matrix, thereby rendering it highly susceptible to contraction-induced injuries that can trigger muscle necrosis (DelloRusso et al. 2001). A growing body of evidence has shown that the DGC is also a transmembrane signaling complex. Therefore, muscle cell death is likely related to disruption of cell survival pathways and cellular defense mechanisms that are regulated by signaling cascades (Lapidos et al. 2004; Muntoni et al. 2003; Thomas et al. 1998).

Animal Models of DMD

There are two naturally occurring animal models for DMD, X-linked mdx mice and X-linked muscular dystrophy dogs (cxmd1). The mdx mouse carries a single-point mutation in exon 23 of the dystrophin gene that results in a premature stop codon (Sicinski et al. 1989). Despite the absence of dystrophin expression in muscle, young mdx mice display a very mild phenotype, apart from the diaphragm, compared to DMD patients, especially with respect to cardiomyopathy. However, as the mice age the phenotype progressively worsens and they display a 20% reduction in lifespan (Chamberlain et al. 2007; Lynch et al. 2001). In addition to the original mdx mouse, researchers have characterized several additional strains that have different mutations and differential expression of dystrophin isoforms with similar pathological phenotype (Cox et al. 1993b; Im et al. 1996). Canine cxmd results from a point mutation in a consensus splice acceptor site in intron 6 of the dystrophin gene, which leads to skipping of exon 7, a disruption in the open reading frame, and premature termination of translation (Howell et al. 1997; Sharp et al. 1992). In contrast to mdx mice, the clinical course of cxmd dogs is very similar to that of DMD humans, characterized by progressive muscle wasting, degeneration and fibrosis, and a shortened life span. But because of the mdx mouse’s low cost and short gestation times , it remains the most widely used animal model (Allamand and Campbell 2000; Gregorevic et al. 2008; Sicinski et al. 1989; Stedman et al. 1991).

Investigators first identified and characterized canine X-linked muscular dystrophy (cxmd) in golden retrievers (Cooper et al. 1988; Kornegay et al. 1988, 1990; Valentine et al. 1988) and then in other breeds including rottweilers (Collins and Morgan 2003) and German shorthaired pointers (Schatzberg et al. 1999). One group produced a beagle model by crossing the golden retriever mutant with beagles (Shimatsu et al. 2003). Litters of golden retrievers crossed with beagles or mongrels have been raised at the Fred Hutchinson Cancer Research Center (Dell’Agnola et al. 2004).

Muscle lesions in cxmd start to develop in utero. Between 6 and 8 weeks of age affected pups begin to show clinical symptoms, which can be quite pronounced by 6 months. The dogs typically die from cardiac or respiratory malfunctions within days, months, or 2 to 4 years after birth (Howell et al. 1997; Valentine et al. 1986, 1988, 1992; Valentine and Cooper 1991).

Because cxmd dogs require extra daily care for reasonable weight and health, the maintenance of a cxmd dog colony is challenging and expensive. However, their severe disease manifestations and progression make them the most useful preclinical animal model for testing therapeutic interventions that have promise for human DMD (Collins and Morgan 2003; Cooper et al. 1988; Howell et al. 1997).

Treatments

Conventional Treatment for DMD

In the absence of curative therapies for human DMD, therapeutic interventions control secondary symptoms with the aim of slowing progression of the disease and improving quality of life.

Although some of these treatments have resulted in improvements of muscle function and a slowing of disease progression in both the mouse model and human patients, adverse side effects have limited their usefulness. Hope for eventual correction of DMD disease symptoms rests with the stable, systemic introduction of a functional dystrophin gene into the muscles of DMD patients.

Developing New Strategies for Treating Dystrophin Deficiency

A number of studies have examined transplantations of normal stem cells such as hematopoietic stem cells (HSCs1), myoblasts, and stem cells derived from other tissue types as possible treatments for DMD and as a gene delivery system for therapeutic recombinant proteins (Camargo et al. 2003; Huard et al. 2003; Lee-Pullen et al. 2004; Parker et al. 2008; Partridge et al. 1998; Torrente et al. 2004). The disease phenotype of the mdx mouse model also shows improvement after the injection, either intramuscular or intravenous (Gregorevic et al. 2004; Gregorevic and Chamberlain 2003; Liu et al. 2005), of viral vectors encoding full-length or truncated but functional dystrophins (DelloRusso et al. 2002; Dudley et al. 2004; Harper et al. 2002b). Furthermore, it is possible to modify the mutant dystrophin mRNA by inducing skipping of the mutation-containing exons, through the use of specific antisense oligonucleotides, sometimes embedded in small nuclear ribonucleoproteins (snRNPs). This strategy can lead to restoration of an open reading frame, which has raised another possible approach for treatment (Alter et al. 2006; Bertoni 2008; Dunckley et al. 1998; McClorey et al. 2006).

However, these approaches are not without many significant hurdles, such as

  • poor survival and limited dissemination of injected cells,

  • immune responses to allogeneic cells, viral vectors, and neotransgene product (especially considering the preexisting inflammatory lesions in dystrophic muscle),

  • unwanted ectopic gene expression, and

  • inability to achieve body-wide muscle delivery (including cardiac tissue) of cells, vectors, and antisense oligonucleotides.

Stem Cells

Stem cells can serve as vehicles to deliver normal genes to mutant organisms. In the case of DMD, it is the hope that stem cells or their progeny carrying a wild-type dystrophin gene would home to the dystrophic muscle, proliferate, and differentiate to form new muscle fibers or fuse to existing myofibers, restoring the missing dystrophin, leading to assembly of the DGC, and eventually ameliorating muscle pathology and improving muscle function.

Hematopoietic stem cells

HSCs include multipotent cells that predominantly reside in the bone marrow and have the capacity for self-renewal and multilineage differentiation (Bellantuono 2004; Lakshmipathy and Verfaillie 2005).2 Hematopoietic cell transplantation (HCT) is a curative treatment for patients with hematological disorders (Storb 2003), but not, so far, for muscular dystrophy. Transplantation of wild-type bone marrow cells into mouse muscles with chemically or crush-induced injury or in the mdx mouse resulted in detectable donor-derived cells and dystrophin-positive fibers in skeletal and cardiac muscle in earlier studies (Bittner et al. 1999; Ferrari et al. 1998; Gussoni et al. 1999) but not in more recent ones (Kucia et al. 2005; LaBarge and Blau 2002; Torrente et al. 2004). Studies in seven cxmd dogs failed to demonstrate a contribution of HSC to skeletal muscle regeneration (Dell’Agnola et al. 2004), despite stable, complete, or near complete donor hematopoietic chimerism. There were no detectable contributions of bone marrow–derived cells to either skeletal muscle (assayed by reverse transcriptase polymerase chain reaction [RT-PCR] and immunoassays) or muscle satellite (cells assayed by clonal analyses). However, muscle cells from the HSC donors did survive (Dell’Agnola et al. 2004). Results from this clinically relevant dog model indicated that allogeneic HCT was of no therapeutic value. Moreover, injections of donor bone marrow cells in the muscles of cxmd dogs made tolerant by preceding DLA-identical HCT also failed to restore dystrophin expression or result in muscle regeneration (Kuhr et al. 2007a).

Muscle stem cells

Muscle-derived stem cells from healthy donors are of potential use for muscle regeneration in DMD patients. The ability of adult skeletal muscle to repair and regenerate itself after injury is largely attributable to muscle satellite cells, which represent a distinct population of myogenic precursors with self-renewal and muscle regeneration properties (Grounds and Davies 2007; Jankowski et al. 2002; Pirenne and Kawai 2004). In DMD, multiple cycles of muscle degeneration and regeneration are associated with a decreasing ability of the satellite cell to efficiently support muscle regeneration (Dhawan and Rando 2005); transplantation of normal myoblasts into dystrophin-deficient muscle may establish a normal myoblast reservoir for delivering dystrophin and regenerating muscle fibers. Fusion of mononucleated myoblasts either into multinucleated muscle fibers or with existing myofibers is a normal part of muscle repair. Normal myoblasts are able to fuse with dystrophic myoblasts to form myotubes that express dystrophin in vitro (Huard et al. 2003; Lee-Pullen et al. 2004). Similarly, intramuscular injection of myoblasts from normal donor mice reconstituted muscle fibers and led to dystrophin expression in mdx mice despite the rapid death of a high percentage of donor cells (Beauchamp et al. 1999; Jejurikar and Kuzon 2003; Lee-Pullen et al. 2004; Skuk et al. 2004). However, myoblast transplantation in DMD patients has been disappointing, with poor survival of injected cells, poor migration of newly introduced myoblasts to damaged areas, and immune responses to both the allogeneic donor cells and wild-type dystrophin (Gussoni et al. 1992; Huard et al. 1991, 1992; Mendell et al. 1995; Partridge 2000; Tremblay et al. 1993). Early studies of myoblast transplantation in dogs from the Tremblay group (Ito et al. 1998) suggested that the combination of three immunosuppressive drugs (FK506; cyclosporine, CSP; and RS-61443, now mycophenolate mofetil, or MMF) effectively controlled immune responses, compared to FK506 alone or CSP plus RS-61443, and allowed engraftment of allogeneic myoblasts for 4 weeks. However, FK506 and CSP-related toxicities were prominent in these dogs. An alternative way to induce tolerance to allogeneic donor myoblasts is to create mixed or complete donor hematopoietic chimerism in recipient dogs treated with DLA-identical HCT (Storb 2003; Storb et al. 1997). This strategy has successfully induced tolerance to donor-derived solid organ transplantation (e.g., kidney, liver and pancreatic islet cells) in mice and dogs without the need for long-term conventional immunosuppression (Kuhr et al. 2007b). A recent study using cxmd dogs (Parker et al. 2008) demonstrated that DLA-identical HCT provided an immune-tolerant platform for subsequent transplantation and stable engraftment of HSC donor-derived myoblasts in the absence of pharmacological immunosuppression. Two chimeric dystrophic dogs received intramuscular injections of freshly isolated myoblasts. The level of wild-type dystrophin mRNA was increased to 6.5% of normal levels in one dog 10 weeks later, and to 1.3% in the other 24 weeks later, and correlated with increased dystrophin protein expression. It is thus possible to induce allogeneic tolerance to donor-specific myoblasts by allogeneic HCT, an option that raises hopes for future use of muscle stem cells as a means of delivering dystrophin to dystrophic muscle.

Mesoangioblasts

After researchers isolated mesoangioblasts from the dorsal aorta of mouse embryos, they found that these blood vessel–derived stem cells can differentiate into various tissue types, including skeletal and cardiac muscles (Cossu and Bianco 2003; De Angelis et al. 1999; Sampaolesi et al. 2003, 2006), and can cross the vascular barrier. Systemic delivery of donor mesoangioblasts via intra-arterial injection in dystrophic and mdx/utrophin null mice led to amelioration of muscle structure and function (Berry et al. 2007; Sampaolesi et al. 2003). Cossu and colleagues tested this approach in dystrophic dogs (Sampaolesi et al. 2006) by isolating wild-type or dystrophic mesoangioblasts from small blood vessels in the muscles of young dogs. They genetically modified dystrophic cells by transduction with a human micro-dystrophin using lentiviral vectors before injection into the femoral artery of the dystrophic dog of origin, and injected allogeneic cells under immunosuppression with CSP or rapamycin into the wild-type dog. All dogs also received steroids. Expression of dystrophin was detected in up to 70% of muscle fibers in the dogs that received wild-type mesoangioblasts, with improved muscle force generation and mobility. The results were encouraging, but critics emphasized the need to clarify several questions. For example, some studies in mdx mice and in DMD patients have suggested some benefit from CSP in ameliorating the dystrophic pathology and improving muscle strength (Davies and Grounds 2006; De Luca et al. 2005; Miller et al. 1997), but no dystrophic dogs that received the same drugs but without mesoangioblast transplantation were used in this study to exclude any effect of immunosuppression. Other questions concern the lack of a strong correlation between muscle function and the extent of expression of dystrophin, and the poor performance of genetically modified autologous cells compared to allogeneic cells despite production of human dystrophin in the canine muscle (Chamberlain 2006; Davies and Grounds 2006; Grounds and Davies 2007).

Viral–Vector Mediated Gene Therapy

The single gene mutation underlying DMD makes viral vector-mediated gene replacement an attractive strategy. Studies have focused on adenoviral (Ad), retroviral, and adeno-associated viral (AAV1) vectors and have shown that delivery via these vectors of full-length or truncated but functional dystrophin improves the disease phenotype of the mdx mouse model (Dudley et al. 2004; Gregorevic et al. 2004; Gregorevic and Chamberlain 2003; Liu et al. 2005). The level of dystrophin expression required for amelioration of skeletal muscle function in mdx mice may be above 20% of normal endogenous levels, with no toxicity seen from a fifty fold over-expression (Cox et al. 1993a; Phelps et al. 1995; Wells et al. 1995), but the minimum and maximum levels of dystrophin expression in cardiac muscle remain to be determined (Duan 2006). Of the vectors used in these studies, Ad and AAV were tested in dog models.

Adenoviral vectors are among the most commonly used vectors for gene therapy (Ghosh et al. 2006) because of their large cloning capacity, non-oncogenic nature, and ability to transduce both dividing and non-dividing cells with high efficiency, including in heart and skeletal muscles. However, a major problem associated with early generations of adenoviral vectors has been robust immune reactions against both viral and transgene products in mdx mice and cxmd golden retrievers (Ghosh et al. 2006; Howell et al. 1998; Kay et al. 2001). “Gutted” adenoviral vectors lack all viral coding sequences and have a cloning capacity of about 35 kb, which is ideal for carrying large genes such as the full-length dystrophin gene (Hartigan-O’Connor et al. 2002). A gutted vector carrying full-length dystrophin can transduce muscle in mdx mice and cxmd golden retrievers, and result in expression of the dystrophin protein in muscle fibers despite humoral immune responses against dystrophin (DelloRusso et al. 2002; Dudley et al. 2004; Gilbert et al. 2001). However, gutted versions of Ad vectors are difficult to grow and scale up to pharmaceutical levels, humoral and cellular reactions in the hosts may result from the use of high vector doses, and intravenous administration of adenoviral vectors results in preferential localization to the liver (Chamberlain 2002; Ghosh et al. 2006; Schiedner et al. 1998). Moreover, Ad vectors do not integrate into the host genome for transgene expression; the relatively short-term expression of the transgene product requires repeated administration.

Retroviral vectors are attractive for treatment of genetic diseases when stable long-term integration in the genome is required. Retroviral vectors can hold up to 11 kb transgene cassettes and show a very low toxicity profile (Chamberlain 2002; Kay et al. 2001; Sinn et al. 2005). Oncoretroviral vectors require dividing cells for efficient cell entry, and transgene expression is dependent on the site of integration and cell division. Moreover, retroviral vectors integrate in somewhat random locations and thus can cause insertional mutagenesis and activation of nearby genes, including oncogenes (Baum et al. 2006). The development of leukemia in part due to ectopic activation of the Lmo2 oncogene in a clinical trial of gene therapy for severe combined immune deficiency disease has confirmed concerns about retroviral vector–based gene delivery (Hacein-Bey-Abina et al. 2002, 2003). Furthermore, the transduction efficiency of retroviral vectors carrying truncated dystrophin into the muscles of mdx mice has been poor and results in only minimal dystrophin expression (Chamberlain 2002; Kay et al. 2001).

Unlike the oncoretroviral vectors, lentiviral vectors based on the human immunodeficiency virus can carry 7.5 to 9 kb cargo DNA, transduce dividing and nondividing cells, and lead to stable expression of transgenes in muscle cells and muscle stem cells (Bachrach et al. 2004; Cudre-Mauroux et al. 2003; Kafri et al. 1997; MacKenzie et al. 2002)

The most promising vector for gene replacement has been AAV, a single-stranded DNA virus belonging to the Parvovirus family with more than 11 serotypes identified in primates (Gao et al. 2004, 2005; Rutledge et al. 1998). AAVs are advantageous for human muscle gene therapy because they transduce non-replicating as well as replicating cells (myofiber and cardiomyocytes are postmitotic), and some serotypes (e.g., AAV1, 6, 8, and 9) exhibit high efficiency for the targeting of striated muscles (Athanasopoulos et al. 2004; Blankinship et al. 2006; Pacak et al. 2006; Wang et al. 2005; Warrington and Herzog 2006). There are no known pathogens associated with AAV in humans, and limited cellular immune responses have been reported after AAV-mediated gene delivery in mice. AAVs are small and can be produced at high titers for robust transduction of muscle through either direct injection or systemic delivery.

The use of AAV vectors as a gene delivery vehicle has shown promise both in preclinical studies and clinical trials for a number of acquired and inherited diseases (Athanasopoulos et al. 2004; Warrington and Herzog 2006). Stable local transgene expression has persisted for years in mice, large animals, and humans (Arruda et al. 2005; Jiang et al. 2006), but AAV vectors have a limited cloning capacity of less than 5 kb. In order to overcome the packaging limitation of AAV for treating DMD, a series of mini- and microdystrophin gene expression cassettes have been developed (Figure 1) based on information from studies of genotype-phenotype relationships in DMD and BMD patients and from transgenic studies in mdx mice (Harper et al. 2002a; Scott et al. 2002). While DMD is caused by frame shift mutations in the dystrophin gene, the milder BMD is typically caused by in-frame mutations in the dystrophin gene. The finding that some BMD patients with large deletions in the central rod domain display milder phenotypes than DMD patients suggested that this domain is of limited function and largely dispensable (Aartsma-Rus et al. 2006; Aartsma-Rus and van Ommen 2007; Abmayr and Chamberlain 2006; England et al. 1990). One of the smallest constructs, Δ R4-R23/ΔCT, is only 3.6 kb and contains the N-terminal domain, the first three and the last one of the 24 spectrin-like repeats and the cysteine-rich domain (Figure 1) (Harper et al. 2002a).

Figure 1. Full-length, mini-, and microdystrophin.

Figure 1

A. Full-length dystrophin; B. minidystrophins with the deletion of various numbers of spectrin repeats; C, microdystrophins with the deletion of various numbers of spectrin repeats with or without C terminus. CR, cysteine-rich domain; CT, carboxyl terminus; R, spectrin repeat; Δ, deletion.

Administration of AAV vectors containing these engineered constructs via either intramuscular or intravenous routes into mdx mice significantly improves muscle membrane integrity and muscle function (Gregorevic et al. 2004; Gregorevic and Chamberlain 2003; Harper et al. 2002a; Liu et al. 2005). The efficacy of this treatment was further evaluated in a clinically relevant cxmd model (Wang et al. 2007a). In sharp contrast to the murine studies, local intramuscular injection of AAV2 or AAV6 vectors into muscles of wild-type dogs induces robust T–cell mediated immune responses against AAV capsid proteins that peaked 4 weeks after vector injection and that limited long-term transgene expression. The immune responses are triggered regardless of the nature of transgene constructs, vector purification methods, or concentrations (from 5 × 109 to 5 × 1012 vector genomes/dog). A follow-up study (Wang et al. 2007b) addressed this issue in both wild-type and dystrophic dogs by evaluating immunosuppression regimens. The study demonstrated that a brief course of immunosuppression with a combination of CSP, MMF, and rabbit anticanine thymocyte globulin (ATG) was able to avert cellular immune responses against AAV6 vectors carrying a canine microdystrophin, allowing sustained dystrophin expression in cxmd muscle, which was sufficient for restoring DGC assembly at the muscle membrane. This observation in a large, immunocompetent, random-bred dog model using a species-specific transgene has set the stage for further investigations of the specificity of immune responses to AAV vectors, the use of alternative, less toxic immunosuppressive regimens, and evaluation of the efficacy and efficiency of systemic ways of delivering transgenes to muscles (Arruda et al. 2005; Su et al. 2005) before embarking on human trials.

Another major challenge is how to achieve bodywide muscle transduction of therapeutic vectors. Systemic gene delivery via intravenous or intra-arterial injection of several AAV serotypes (AAV6, 8, and 9) in rodents demonstrated robust transduction throughout the body, including cardiac muscle (Gregorevic et al. 2004; Inagaki et al. 2006; Pacak et al. 2006; Wang et al. 2006). However, there is a significant difference in body size between rodents and large vertebrates, and the differences in anatomy and physiology, including vascular permeability, also pose great challenges to translation of this technique to humans. A recent study from Duan’s group suggested that bodywide delivery of therapeutic genes may be an achievable goal in larger animals (Yue et al. 2008). In that study, four neonatal dogs (24 or 48 hours after birth) received single injections in the jugular vein of AAV9 vectors at doses ranging from 1 × 1011 to 2.5 × 1011 vg/gbody weight. Robust whole-body skeletal muscle transduction occurred in all cases, and expression of the transgenes persisted for 6 months without detectable cellular infiltrations. One of the four pups contracted a deadly illness after vector administration and the authors suggested this might have been due to a higher vector dose (2.5 × 1011 vg/g) given at an early age (24 hours after birth). Otherwise, the study results certainly raised hopes for potential application of systemic AAV delivery in human patients. However, one has to keep in mind when considering treatment for adult humans that the immune system in newborns is immature compared to that of adults, and vessel permeability may be different in young versus adult animals. Cardiomyopathy is a leading cause of death in DMD patients, and no cardiac muscle transduction was observed in this study, in contrast to murine data. The safety of widespread dissemination of virus throughout the body as a result of systemic delivery should also be taken into consideration when adapting these strategies to clinical trials.

Antisense Oligonucleotide-Mediated Gene Correction

Another strategy to convert a severe DMD phenotype to a milder BMD phenotype is to modulate splicing of the dystrophin mRNA and skip exons containing mutations that disrupt the open reading frame (Aartsma-Rus et al. 2003; Muntoni et al. 2003). Studies of targeted exon skipping have used several types of antisense oligonucleotides (AO), which include 2′-O-methyl phosphorothioate (2OMe) and modified alternatives, phosphorodiamidate morpholino oligomers (PMOs), and peptide nucleic acid (PNA) AOs. These studies are aimed at modulating gene expression rather than adding new genes (reviewed in Aartsma-Rus and van Ommen 2007; Foster et al. 2006). The mechanism of exon skipping is based on the binding of AOs to specific target sense sequences in the dystrophin pre-mRNA. This binding may alter the local configuration of the pre-mRNA so the splicing machinery can no longer recognize it and therefore skips it, or interfere with components of the splicing machinery (Aartsma-Rus et al. 2003; van Deutekom and van Ommen 2003). The end result is to remove a portion of the mRNA and restore an open reading frame. Theoretical predictions have suggested that targeting one or more of only twelve exons in the dystrophin gene could lead to treatment options for approximately 75% of the mutations that cause human DMD (Aartsma-Rus et al. 2003; Aartsma-Rus and van Ommen 2007; Bertoni 2008; van Deutekom and van Ommen 2003). In vitro studies using human and mouse muscle cells (Aartsma-Rus et al. 2002, 2003) and in vivo studies using the mdx mouse (Alter et al. 2006; Dunckley et al. 1998; Vitiello et al. 2008) have shown promising results in restoring the open reading frame, thereby leading to expression of smaller, truncated, but functional dystrophins. A recent study from the Wilton group (McClorey et al. 2006) addressed the feasibility of this approach in the canine DMD model in vitro. The study evaluated the efficacy of 2OMe, PMOs, and peptide-linked PMOs to induce truncated in-frame dystrophin expression with the skipping of both exons 6 and 8 to restore the disrupted reading frame due to loss of exon 7 in cultured myoblasts isolated from golden retriever cxmd dogs. The group observed that 2OMe AOs were short-lived and caused moderate cell death at a low concentration of 600 nM, whereas PMO and peptide-linked PMO did not cause obvious toxicity to cells at a concentration of 20 μM. However, unconjugated PMO alone transduced cells poorly with limited expression of dystrophin induced at high concentration. In contrast, conjugated transport peptide facilitated transfection of PMOs and induced high levels of exon skipping for at least 10 days with high levels of dystrophin expression.

Whether or not antisense treatment can result in functional benefit remains to be tested in vivo in dogs with DMD. AOs are small, sequence-specific, and synthetic, and are considered relatively safe. They also have the advantage of being able to simultaneously correct all dystrophin isoforms (Aartsma-Rus et al. 2003; Wilton et al. 1999). But there remain several obstacles associated with the effectiveness of the treatment. AOs do not transduce cells efficiently, they can easily be degraded (which limits the duration of effectiveness), and they require repeated administration, probably weekly or monthly for the life of the patient. AOs linked to a modified U7 small nuclear RNA, normally involved in mRNA processing, achieved sustained dystrophin expression from skipped mRNA for more than 13 weeks in the limbs of the mdx mouse (Goyenvalle et al. 2004), thereby raising the hope for potential clinical application of this strategy. To date, AO treatment has not successfully induced dystrophin expression in cardiac muscle. In addition, the approach has the potential risk of causing nonspecific splicing aberrations at high concentration, and efficient systemic delivery methods will be required for muscle-specific targeting. Nevertheless, antisense therapy remains one of the more promising strategies for the treatment of most DMD patients.

Perspectives

The discovery and development of new therapy options for muscular dystrophies are ongoing. For stem cell–based approaches, investigators isolated a distinct subpopulation of muscle satellite cells in mice using a combination of markers including CXCR4 and β1-integrin (Cerletti et al. 2008). After intramuscular injection, these cells entered and renewed the endogenous satellite cell pool, participated in repairing damaged muscle by both fusing to existing muscle fibers and forming new fibers, and contributed to more than 90% of myofibers in injected muscle. There is growing interest in using nonviral transfer methods for delivering therapeutic genes, such as the use of naked plasmid DNA to deliver dystrophin cDNA constructs (reviewed in Rando 2007; Scime and Rudnicki 2008); intravascular injection of naked DNA into rodents and pigs showed encouraging results (Danialou et al. 2005; Wolff et al. 2005). PTC124, a newly identified chemical entity, promotes read-through of premature nonsense codon as well as the production of dystrophin protein in muscle cells derived from mdx mice and human patients with dystrophin genes containing nonsense mutations (Welch et al. 2007). Compensatory therapeutic strategies include the possibilities of promoting muscle regeneration and ameliorating the disease phenotype either (1) to increase the expression of utrophin (a homologue of dystrophin) to compensate for impaired dystrophin function or (2) to inhibit myostatin (a negative regulator of muscle growth) and thus increase muscle mass (Bogdanovich et al. 2002; Fisher et al. 2001; Qiao et al. 2008a,b; Wakefield et al. 2000).

The advances in the last two decades in the field of DMD therapeutics are remarkable and have energized the planning and implementation of phase I clinical trials. For example, high-density intramuscular injections of muscle stem cells in DMD patients are the subject of tests by the Tremblay group (Skuk et al. 2006), antisense oligonucleotides targeting exon 51 in DMD patients are being tested by van Deutekom and colleagues (2007), naked plasmids carrying the full-length dystrophin cDNA under the control of the cytomegalovirus (CMV) promoter have been injected into DMD and BMD boys (Romero et al. 2004), and the intramuscular injection of AAV vectors carrying functional human microdystrophin in DMD patients is under investigation by the Mendel group (Rodino-Klapac et al. 2007). The results from the studies that have been made public suggest the feasibility of these approaches in humans, but considerable further investigation is necessary to overcome hurdles such as immune responses, systemic delivery efficiencies, and long-term dystrophin expression. Ideally, such studies will be optimized in large animal models before a successful clinical translation. Ultimately, effective treatment may require combinations of several of these approaches.

Acknowledgments

This work is supported by NIH U54-HD47175 and the Seattle Senator Paul D. Wellstone Muscular Dystrophy Cooperative Research Center, NIH CA15704, and funds from the Muscular Dystrophy Association and the Parent Project.

Glossary

AAV

adeno-associated virus, a single-stranded DNA virus belonging to the Parvovirus family. Widely used as a gene transfer vehicle in the gene therapy field

AO

antisense oligonucleotides, used for targeted exon skipping. AOs can bind to specific target sense sequences within pre-mRNA to alter the local configuration of the pre-mRNA either such that the splicing machinery can no longer recognize it and therefore skips it, or by interfering with components of the splicing machinery

BMD

Becker muscular dystrophy, a milder form of DMD that results from truncated yet partly functional dystrophin protein as a result of in-frame mutations in the dystrophin gene

Cxmd

canine X-linked muscular dystrophy, caused by dystrophin deficiency due to frame shift point mutation in the canine dystrophin gene. It has clinical and pathological courses that are very similar to human DMD

DGC

dystrophin-glycoprotein complex, which spans the plasma membrane and provides a strong mechanical link connecting the intracellular γ-actin cytoskeleton to the extracellular matrix at the sarcolemma of striated and smooth muscles

DMD

Duchenne muscular dystrophy, a lethal X-linked form of muscular dystrophy caused by the loss of the protein, dystrophin, a critical component of the dystrophin-glycoprotein complex (DGC) at the muscle membrane

HCT

hematopoietic cell transplantation, used as curative treatment for patients with many hematological disorders

HSC

hematopoietic stem cells, multipotent cells that predominantly reside in the bone marrow and have the capacity for self-renewal and multilineage differentiation

Footnotes

1

Abbreviations used in this article: AAV, adeno-associated virus; AO, antisense oligonucleotides; BMD, Becker muscular dystrophy; cxmd, canine X-linked muscular dystrophy; etc.

2

Multipotent cells also give rise to nonhematopoietic cells, as in bone and liver, vascular endothelial cells, and astroglia in the brain (Bhattacharya et al. 2000; Shi et al. 1998; Torrente et al. 2004), possibly through transdedifferentiation or cell fusion (Bellantuono 2004) or recruitment of subpopulations of tissue-committed stem cells (Lakshmipathy and Verfaillie 2005; Torrente et al. 2004).

Contributor Information

Zejing Wang, Research Associate in the Program in Transplantation Biology of the Division of Clinical Research at the Fred Hutchinson Cancer Research Center in Seattle, Washington.

Jeffrey S. Chamberlain, Professor in the Departments of Neurology, Biochemistry, and Medicine at the University of Washington in Seattle.

Stephen J. Tapscott, Professor in the University of Washington Department of Medicine and Full Member in the Division of Human Biology at the Fred Hutchinson Cancer Research Center.

Rainer Storb, [Full Member in [or of] the Program in Transplantation Biology of the Division of Clinical Research at the Fred Hutchinson Cancer Research Center and Professor in the Department of Medicine at the University of Washington.

References

  1. Aartsma-Rus A, van Ommen GJ. Antisense-mediated exon skipping: A versatile tool with therapeutic and research applications. RNA. 2007;13:1609–1624. doi: 10.1261/rna.653607. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Aartsma-Rus A, Bremmer-Bout M, Janson AA, Den Dunnen JT, van Ommen GJ, van Deutekom JC. Targeted exon skipping as a potential gene correction therapy for Duchenne muscular dystrophy. Neuromusc Dis. 2002;12 (Suppl 1):S71–S77. doi: 10.1016/s0960-8966(02)00086-x. [DOI] [PubMed] [Google Scholar]
  3. Aartsma-Rus A, Janson AA, Kaman WE, Bremmer-Bout M, Den Dunnen JT, Baas F, van Ommen GJ, van Deutekom JC. Therapeutic antisense-induced exon skipping in cultured muscle cells from six different DMD patients. Hum Mol Genet. 2003;12:907–914. doi: 10.1093/hmg/ddg100. [DOI] [PubMed] [Google Scholar]
  4. Aartsma-Rus A, van Deutekom JC, Fokkema IF, van Ommen GJ, Den Dunnen JT. Entries in the Leiden Duchenne muscular dystrophy mutation database: An overview of mutation types and paradoxical cases that confirm the reading-frame rule. Musc Nerve. 2006;34:135–144. doi: 10.1002/mus.20586. [DOI] [PubMed] [Google Scholar]
  5. Abmayr S, Chamberlain J. The structure and function of dystrophin. In: Winder SJ, editor. Molecular Mechanisms of Muscular Dystrophies. Georgetown TX: Landes Bioscience; 2006. pp. 14–47. [Google Scholar]
  6. Allamand V, Campbell KP. Animal models for muscular dystrophy: Valuable tools for the development of therapies. Hum Mol Genet. 2000;9:2459–2467. doi: 10.1093/hmg/9.16.2459. [DOI] [PubMed] [Google Scholar]
  7. Alter J, Lou F, Rabinowitz A, Yin H, Rosenfeld J, Wilton SD, Partridge TA, Lu QL. Systemic delivery of morpholino oligonucleotide restores dystrophin expression bodywide and improves dystrophic pathology. Nat Med. 2006;12:175–177. doi: 10.1038/nm1345. [DOI] [PubMed] [Google Scholar]
  8. Arruda VR, Stedman HH, Nichols TC, Haskins ME, Nicholson M, Herzog RW, Couto LB, High KA. Regional intravascular delivery of AAV-2-F.IX to skeletal muscle achieves long-term correction of hemophilia B in a large animal model. Blood. 2005;105:3458–3464. doi: 10.1182/blood-2004-07-2908. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Athanasopoulos T, Graham IR, Foster H, Dickson G. Recombinant adeno-associated viral (rAAV) vectors as therapeutic tools for Duchenne muscular dystrophy (DMD) Gene Ther. 2004;11(Suppl 1):S109–S121. doi: 10.1038/sj.gt.3302379. [DOI] [PubMed] [Google Scholar]
  10. Bachrach E, Li S, Perez AL, Schienda J, Liadaki K, Volinski J, Flint A, Chamberlain J, Kunkel LM. Systemic delivery of human microdystrophin to regenerating mouse dystrophic muscle by muscle progenitor cells. Proc Natl Acad Sci U S A. 2004;101:3581–3586. doi: 10.1073/pnas.0400373101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Baum C, Kustikova O, Modlich U, Li Z, Fehse B, Authors FN. Mutagenesis and oncogenesis by chromosomal insertion of gene transfer vectors. Hum Gene Ther. 2006;17:253–263. doi: 10.1089/hum.2006.17.253. [DOI] [PubMed] [Google Scholar]
  12. Beauchamp JR, Morgan JE, Pagel CN, Partridge TA. Dynamics of myoblast transplantation reveal a discrete minority of precursors with stem cell-like properties as the myogenic source. J Cell Biol. 1999;144:1113–1122. doi: 10.1083/jcb.144.6.1113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Bellantuono I. Hæmopoietic stem cells. Int J Biochem Cell B. 2004;36:607–620. doi: 10.1016/j.biocel.2003.10.008. [DOI] [PubMed] [Google Scholar]
  14. Berry SE, Liu J, Chaney EJ, Kaufman SJ. Multipotential mesoangioblast stem cell therapy in the mdx/utrn-/-mouse model for Duchenne muscular dystrophy. Regen Med. 2007;2:275–288. doi: 10.2217/17460751.2.3.275. [DOI] [PubMed] [Google Scholar]
  15. Bertoni C. Clinical approaches in the treatment of Duchenne muscular dystrophy (DMD) using oligonucleotides. Front Biosci. 2008;13:517–527. doi: 10.2741/2697. [DOI] [PubMed] [Google Scholar]
  16. Bittner RE, Schofer C, Weipoltshammer K, Ivanova S, Streubel B, Hauser E, Freilinger M, Hoger H, Elbe-Burger A, Wachtler F. Recruitment of bone-marrow-derived cells by skeletal and cardiac muscle in adult dystrophic mdx mice. Anat Embryol. 1999;199:391–396. doi: 10.1007/s004290050237. [DOI] [PubMed] [Google Scholar]
  17. Blankinship MJ, Gregorevic P, Chamberlain JS. Gene therapy strategies for Duchenne muscular dystrophy utilizing recombinant adeno-associated virus vectors. Mol Ther. 2006;13:241–249. doi: 10.1016/j.ymthe.2005.11.001. [DOI] [PubMed] [Google Scholar]
  18. Bogdanovich S, Krag TO, Barton ER, Morris LD, Whittemore LA, Ahima RS, Khurana TS. Functional improvement of dystrophic muscle by myostatin blockade. Nature. 2002;420:418–421. doi: 10.1038/nature01154. [DOI] [PubMed] [Google Scholar]
  19. Burdi R, Didonna MP, Pignol B, Nico B, Mangieri D, Rolland JF, Camerino C, Zallone A, Ferro P, Andreetta F, Confalonieri P, De Luca A. First evaluation of the potential effectiveness in muscular dystrophy of a novel chimeric compound, BN 82270, acting as calpain-inhibitor and anti-oxidant. Neuromusc Dis. 2006;16:237–248. doi: 10.1016/j.nmd.2006.01.013. [DOI] [PubMed] [Google Scholar]
  20. Byers TJ, Lidov HG, Kunkel LM. An alternative dystrophin transcript specific to peripheral nerve. Nat Genet. 1993;4:77–81. doi: 10.1038/ng0593-77. [DOI] [PubMed] [Google Scholar]
  21. Camargo FD, Green R, Capetenaki Y, Jackson KA, Goodell MA. Single hematopoietic stem cells generate skeletal muscle through myeloid intermediates [erratum in Nat Med 10:105] Nat Med. 2003;9:1520–1527. doi: 10.1038/nm963. [DOI] [PubMed] [Google Scholar]
  22. Cerletti M, Jurga S, Witczak CA, Hirshman MF, Shadrach JL, Goodyear LJ, Wagers AJ. Highly efficient, functional engraftment of skeletal muscle stem cells in dystrophic muscles. Cell. 2008;134:37–47. doi: 10.1016/j.cell.2008.05.049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Chamberlain JS. Gene therapy of muscular dystrophy. Hum Mol Genet. 2002;11:2355–2362. doi: 10.1093/hmg/11.20.2355. [DOI] [PubMed] [Google Scholar]
  24. Chamberlain JS. Stem-cell biology: A move in the right direction. Nature. 2006;444:552–553. doi: 10.1038/nature05406. [DOI] [PubMed] [Google Scholar]
  25. Chamberlain JS, Metzger J, Reyes M, Townsend D, Faulkner JA. Dystrophin-deficient mdx mice display a reduced life span and are susceptible to spontaneous rhabdomyosarcoma. FASEB J. 2007;21:2195–2204. doi: 10.1096/fj.06-7353com. [DOI] [PubMed] [Google Scholar]
  26. Collins CA, Morgan JE. Duchenne’s muscular dystrophy: Animal models used to investigate pathogenesis and develop therapeutic strategies. Int J Exp Pathol. 2003;84:165–172. doi: 10.1046/j.1365-2613.2003.00354.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Cooper BJ, Winand NJ, Stedman H, Valentine BA, Hoffman EP, Kunkel LM, Scott M-O, Fischbeck KH, Kornegay JN, Avery RJ, Williams JR, Schmickel RD, Sylvester JE. The homologue of the Duchenne locus is defective in X-linked muscular dystrophy of dogs. Nature. 1988;334:154–156. doi: 10.1038/334154a0. [DOI] [PubMed] [Google Scholar]
  28. Cossu G, Bianco P. Mesoangioblasts: Vascular progenitors for extravascular mesodermal tissues. Curr Opin Genet Dev. 2003;13:537–542. doi: 10.1016/j.gde.2003.08.001. [DOI] [PubMed] [Google Scholar]
  29. Cox GA, Cole NM, Matsumura K, Phelps SF, Hauschka SD, Campbell KP, Faulkner JA, Chamberlain JS. Overexpression of dystrophin in transgenic mdx mice eliminates dystrophic symptoms without toxicity. Nature. 1993a;364:725–729. doi: 10.1038/364725a0. [DOI] [PubMed] [Google Scholar]
  30. Cox GA, Phelps SF, Chapman VM, Chamberlain JS. New mdx mutation disrupts expression of muscle and nonmuscle isoforms of dystrophin. Nat Genet. 1993b;4:87–93. doi: 10.1038/ng0593-87. [DOI] [PubMed] [Google Scholar]
  31. Cudre-Mauroux C, Occhiodoro T, Konig S, Salmon P, Bernheim L, Trono D. Lentivector-mediated transfer of Bmi-1 and telomerase in muscle satellite cells yields a Duchenne myoblast cell line with long-term genotypic and phenotypic stability. Hum Gene Ther. 2003;14:1525–1533. doi: 10.1089/104303403322495034. [DOI] [PubMed] [Google Scholar]
  32. Danialou G, Comtois AS, Matecki S, Nalbantoglu J, Karpati G, Gilbert R, Geoffroy P, Gilligan S, Tanguay JF, Petrof BJ. Optimization of regional intraarterial naked DNA-mediated transgene delivery to skeletal muscles in a large animal model. Mol Ther. 2005;11:257–266. doi: 10.1016/j.ymthe.2004.09.016. [DOI] [PubMed] [Google Scholar]
  33. Davies KE, Grounds MD. Treating muscular dystrophy with stem cells? Cell. 2006;127:1304–1306. doi: 10.1016/j.cell.2006.12.010. [DOI] [PubMed] [Google Scholar]
  34. De Angelis L, Berghella L, Coletta M, Lattanzi L, Zanchi M, Cusella-De Angelis MG, Ponzetto C, Cossu G. Skeletal myogenic progenitors originating from embryonic dorsal aorta coexpress endothelial and myogenic markers and contribute to postnatal muscle growth and regeneration. J Cell Biol. 1999;147:869–878. doi: 10.1083/jcb.147.4.869. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. De Luca A, Nico B, Liantonio A, Didonna MP, Fraysse B, Pierno S, Burdi R, Mangieri D, Rolland JF, Camerino C, Zallone A, Confalonieri P, Andreetta F, Arnoldi E, Courdier-Fruh I, Magyar JP, Frigeri A, Pisoni M, Svelto M, Conte CD. A multidisciplinary evaluation of the effectiveness of cyclosporine a in dystrophic mdx mice. Am J Pathol. 2005;166:477–489. doi: 10.1016/S0002-9440(10)62270-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Dell’Agnola C, Wang Z, Storb R, Tapscott SJ, Kuhr CS, Hauschka SD, Lee RS, Sale GE, Zellmer E, Gisburne S, Bogan J, Kornegay JN, Cooper BJ, Gooley TA, Little M-T. Hematopoietic stem cell transplantation does not restore dystrophin expression in Duchenne muscular dystrophy dogs. Blood. 2004;104:4311–4318. doi: 10.1182/blood-2004-06-2247. [DOI] [PubMed] [Google Scholar]
  37. DelloRusso C, Crawford RW, Chamberlain JS, Brooks SV. Tibialis anterior muscles in mdx mice are highly susceptible to contraction-induced injury. J Muscle Res Cell M. 2001;22:467–475. doi: 10.1023/a:1014587918367. [DOI] [PubMed] [Google Scholar]
  38. DelloRusso C, Scott JM, Hartigan-O’Connor D, Salvatori G, Barjot C, Robinson AS, Crawford RW, Brooks SV, Chamberlain JS. Functional correction of adult mdx mouse muscle using gutted adenoviral vectors expressing full-length dystrophin. Proc Natl Acad Sci U S A. 2002;99:12979–12984. doi: 10.1073/pnas.202300099. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Dhawan J, Rando TA. Stem cells in postnatal myogenesis: Molecular mechanisms of satellite cell quiescence, activation and replenishment. Trends Cell Bio. 2005;15:666–673. doi: 10.1016/j.tcb.2005.10.007. [DOI] [PubMed] [Google Scholar]
  40. Duan D. Challenges and opportunities in dystrophin-deficient cardiomyopathy gene therapy. Hum Mol Genet. 2006;15 (Spec2):R253–R261. doi: 10.1093/hmg/ddl180. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Dudley RW, Lu Y, Gilbert R, Matecki S, Nalbantoglu J, Petrof BJ, Karpati G. Sustained improvement of muscle function one year after full-length dystrophin gene transfer into mdx mice by a gutted helper-dependent adenoviral vector. Hum Gene Ther. 2004;15:145–156. doi: 10.1089/104303404772679959. [DOI] [PubMed] [Google Scholar]
  42. Dunckley MG, Manoharan M, Villiet P, Eperon IC, Dickson G. Modification of splicing in the dystrophin gene in cultured mdx muscle cells by antisense oligoribonucleotides. Hum Mol Genet. 1998;7:1083–1090. doi: 10.1093/hmg/7.7.1083. [DOI] [PubMed] [Google Scholar]
  43. Emery AE. The muscular dystrophies. Lancet. 2002;359:687–695. doi: 10.1016/S0140-6736(02)07815-7. [DOI] [PubMed] [Google Scholar]
  44. England SB, Nicholson LV, Johnson MA, Forrest SM, Love DR, Zubrzycka-Gaarn EE, Bulman DE, Harris JB, Davies KE. Very mild muscular dystrophy associated with the deletion of 46% of dystrophin. Nature. 1990;343:180–182. doi: 10.1038/343180a0. [DOI] [PubMed] [Google Scholar]
  45. Ervasti JM. Structure and function of the dystrophin-glycoprotein complex. In: Winder SJ, editor. Molecular Mechanisms of Muscular Dystrophies. Georgetown TX: Landes Bioscience; 2006. pp. 1–13. [Google Scholar]
  46. Ferrari G, Cusella-De AG, Coletta M, Paolucci E, Stornaiuolo A, Cossu G, Mavilio F. Muscle regeneration by bone marrow-derived myogenic progenitors. Science. 1998;279:1528–1530. doi: 10.1126/science.279.5356.1528. [DOI] [PubMed] [Google Scholar]
  47. Fisher R, Tinsley JM, Phelps SR, Squire SE, Townsend ER, Martin JE, Davies KE. Non-toxic ubiquitous over-expression of utrophin in the mdx mouse. Neuromusc Dis. 2001;11:713–721. doi: 10.1016/s0960-8966(01)00220-6. [DOI] [PubMed] [Google Scholar]
  48. Foidart M, Foidart JM, Engel WK. Collagen localization in normal and fibrotic human skeletal muscle. Arch Neurol. 1981;38:152–157. doi: 10.1001/archneur.1981.00510030046006. [DOI] [PubMed] [Google Scholar]
  49. Foster K, Foster H, Dickson JG. Gene therapy progress and prospects: Duchenne muscular dystrophy. Gene Ther. 2006;13:1677–1685. doi: 10.1038/sj.gt.3302877. [DOI] [PubMed] [Google Scholar]
  50. Gao G, Vandenberghe LH, Alvira MR, Lu Y, Calcedo R, Zhou X, Wilson JM. Clades of adeno-associated viruses are widely disseminated in human tissues. J Virol. 2004;78:6381–6388. doi: 10.1128/JVI.78.12.6381-6388.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Gao G, Vandenberghe LH, Wilson JM. New recombinant serotypes of AAV vectors. Curr Gene Ther. 2005;5:285–297. doi: 10.2174/1566523054065057. [DOI] [PubMed] [Google Scholar]
  52. Ghosh SS, Gopinath P, Ramesh A. Adenoviral vectors: A promising tool for gene therapy. Appl Biochem Biotech. 2006;133:9–29. doi: 10.1385/abab:133:1:9. [DOI] [PubMed] [Google Scholar]
  53. Gilbert R, Nalbantoglu J, Howell JM, Davies L, Fletcher S, Amalfitano A, Petrof BJ, Kamen A, Massie B, Karpati G. Dystrophin expression in muscle following gene transfer with a fully deleted (“gutted”) adenovirus is markedly improved by trans-acting adenoviral gene products. Hum Gene Ther. 2001;12:1741–1755. doi: 10.1089/104303401750476249. [DOI] [PubMed] [Google Scholar]
  54. Goyenvalle A, Vulin A, Fougerousse F, Leturcq F, Kaplan JC, Garcia L, Danos O. Rescue of dystrophic muscle through U7 snRNA-mediated exon skipping. Science. 2004;306:1796–1799. doi: 10.1126/science.1104297. [DOI] [PubMed] [Google Scholar]
  55. Gregorevic P, Chamberlain JS. Gene therapy for muscular dystrophy: A review of promising progress. Expert Opin Biol Th. 2003;3:803–814. doi: 10.1517/14712598.3.5.803. [DOI] [PubMed] [Google Scholar]
  56. Gregorevic P, Blankinship MJ, Allen JM, Crawford RW, Meuse L, Miller DG, Russell DW, Chamberlain JS. Systemic delivery of genes to striated muscles using adeno-associated viral vectors. Nat Med. 2004;10:828–834. doi: 10.1038/nm1085. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Gregorevic P, Blankinship MJ, Allen JM, Chamberlain JS. Systemic microdystrophin gene delivery improves skeletal muscle structure and function in old dystrophic mdx mice. Mol Ther. 2008;16:657–664. doi: 10.1038/mt.2008.28. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Grounds MD, Davies KE. The allure of stem cell therapy for muscular dystrophy. Neuromusc Dis. 2007;17:206–208. doi: 10.1016/j.nmd.2007.01.007. [DOI] [PubMed] [Google Scholar]
  59. Grounds MD, Torrisi J. Anti-TNFalpha (Remicade) therapy protects dystrophic skeletal muscle from necrosis. FASEB J. 2004;18:676–682. doi: 10.1096/fj.03-1024com. [DOI] [PubMed] [Google Scholar]
  60. Gussoni E, Pavlath GK, Lanctot AM, Sharma KR, Miller RG, Steinman L, Blau HM. Normal dystrophin transcripts detected in Duchenne muscular dystrophy patients after myoblast transplantation. Nature. 1992;356:435–438. doi: 10.1038/356435a0. [DOI] [PubMed] [Google Scholar]
  61. Gussoni E, Soneoka Y, Strickland CD, Buzney EA, Khan MK, Flint AF, Kunkel LM, Mulligan RC. Dystrophin expression in the mdx mouse restored by stem cell transplantation. Nature. 1999;401:390–394. doi: 10.1038/43919. [DOI] [PubMed] [Google Scholar]
  62. Hacein-Bey-Abina S, Le Deist F, Carlier F, Bouneaud C, Hue C, De Villartay JP, Thrasher AJ, Wulffraat N, Sorensen R, Dupuis-Girod S, Fischer A, Davies EG, Kuis W, Leiva L, Cavazzana-Calvo M. Sustained correction of X-linked severe combined immunodeficiency by ex vivo gene therapy. New Engl J Med. 2002;346:1185–1193. doi: 10.1056/NEJMoa012616. [DOI] [PubMed] [Google Scholar]
  63. Hacein-Bey-Abina S, von Kalle C, Schmidt M, McCormack MP, Wulffraat N, Leboulch P, Lim A, Osborne CS, Pawliuk R, Morillon E, Sorensen R, Forster A, Fraser P, Cohen JI, de Saint Basile G, Alexander I, Wintergerst U, Frebourg T, Aurias A, Stoppa-Lyonnet D, Romana S, Radford-Weiss I, Gross F, Valensi F, Delabesse E, Macintyre E, Sigaux F, Soulier J, Leiva LE, Wissler M, Prinz C, Rabbitts TH, Le Deist F, Fischer A, Cavazzana-Calvo M. LMO2-associated clonal T cell proliferation in two patients after gene therapy for SCID-X1 [erratum in Science 302:568] Science. 2003;302:415–419. doi: 10.1126/science.1088547. [DOI] [PubMed] [Google Scholar]
  64. Harper SQ, Crawford RW, DelloRusso C, Chamberlain JS. Spectrin-like repeats from dystrophin and alpha-actinin-2 are not functionally interchangeable. Hum Mol Genet. 2002a;11:1807–1815. doi: 10.1093/hmg/11.16.1807. [DOI] [PubMed] [Google Scholar]
  65. Harper SQ, Hauser MA, DelloRusso C, Duan D, Crawford RW, Phelps SF, Harper HA, Robinson AS, Engelhardt JF, Brooks SV, Chamberlain JS. Modular flexibility of dystrophin: Implications for gene therapy of Duchenne muscular dystrophy. Nat Med. 2002b;8:253–261. doi: 10.1038/nm0302-253. [DOI] [PubMed] [Google Scholar]
  66. Hartigan-O’Connor D, Barjot C, Salvatori G, Chamberlain JS. Generation and growth of gutted adenoviral vectors. Meth Enzymol. 2002;346:224–246. doi: 10.1016/s0076-6879(02)46058-2. [DOI] [PubMed] [Google Scholar]
  67. Hoffman EP, Brown RHJ, Kunkel LM. Dystrophin: The protein product of the Duchenne muscular dystrophy locus. Cell. 1987;51:919–928. doi: 10.1016/0092-8674(87)90579-4. [DOI] [PubMed] [Google Scholar]
  68. Hoffman EP, Kunkel LM, Brown RH., Jr Proteolytic fragment or new gene product? Nature. 1988;336:210. doi: 10.1038/336210a0. [DOI] [PubMed] [Google Scholar]
  69. Howell JM, Fletcher S, Kakulas BA, O’Hara M, Lochmuller H, Karpati G. Use of the dog model for Duchenne muscular dystrophy in gene therapy trials. Neuromusc Dis. 1997;7:325–328. doi: 10.1016/s0960-8966(97)00057-6. [DOI] [PubMed] [Google Scholar]
  70. Howell JM, Lochmüller H, O’Hara A, Fletcher S, Kakulas BA, Massie B, Nalbantoglu J, Karpati G. High-level dystrophin expression after adenovirus-mediated dystrophin minigene transfer to skeletal muscle of dystrophic dogs: Prolongation of expression with immunosuppression. Hum Gene Ther. 1998;9:629–634. doi: 10.1089/hum.1998.9.5-629. [DOI] [PubMed] [Google Scholar]
  71. Huard J, Bouchard JP, Roy R, Labrecque C, Dansereau G, Lemieux B, Tremblay JP. Myoblast transplantation produced dystrophin-positive muscle fibres in a 16-year-old patient with Duchenne muscular dystrophy. Clin Sci. 1991;81:287–288. doi: 10.1042/cs0810287. [DOI] [PubMed] [Google Scholar]
  72. Huard J, Roy R, Bouchard JP, Malouin F, Richards CL, Tremblay JP. Human myoblast transplantation between immunohistocompatible donors and recipients produces immune reactions. Transpl Proc. 1992;24:3049–3051. [PubMed] [Google Scholar]
  73. Huard J, Cao B, Qu-Petersen Z. Muscle-derived stem cells: Potential for muscle regeneration. Birth Defects Res C. 2003;69:230–237. doi: 10.1002/bdrc.10020. [DOI] [PubMed] [Google Scholar]
  74. Im WB, Phelps SF, Copen EH, Adams EG, Slightom JL, Chamberlain JS. Differential expression of dystrophin isoforms in strains of mdx mice with different mutations. Hum Mol Genet. 1996;5:1149–1153. doi: 10.1093/hmg/5.8.1149. [DOI] [PubMed] [Google Scholar]
  75. Inagaki K, Fuess S, Storm TA, Gibson GA, Mctiernan CF, Kay MA, Nakai H. Robust systemic transduction with AAV9 vectors in mice: Efficient global cardiac gene transfer superior to that of AAV8. Mol Ther. 2006;14:45–53. doi: 10.1016/j.ymthe.2006.03.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Ito H, Vilquin JT, Skuk D, Roy B, Goulet M, Lille S, Dugre FJ, Asselin I, Roy R, Fardeau M, Tremblay JP. Myoblast transplantation in non-dystrophic dog. Neuromusc Dis. 1998;8:95–110. doi: 10.1016/s0960-8966(97)00148-x. [DOI] [PubMed] [Google Scholar]
  77. Jankowski RJ, Deasy BM, Huard J. Muscle-derived stem cells. Gene Ther. 2002;9:642–647. doi: 10.1038/sj.gt.3301719. [DOI] [PubMed] [Google Scholar]
  78. Jejurikar SS, Kuzon WM., Jr Satellite cell depletion in degenerative skeletal muscle. Apoptosis. 2003;8:573–578. doi: 10.1023/A:1026127307457. [DOI] [PubMed] [Google Scholar]
  79. Jiang H, Pierce GF, Ozelo MC, de Paula EV, Vargas JA, Smith P, Sommer J, Luk A, Manno CS, High KA, Arruda VR. Evidence of multiyear factor IX expression by AAV-mediated gene transfer to skeletal muscle in an individual with severe hemophilia B. Mol Ther. 2006;14:452–455. doi: 10.1016/j.ymthe.2006.05.004. [DOI] [PubMed] [Google Scholar]
  80. Kafri T, Blomer U, Peterson DA, Gage FH, Verma IM. Sustained expression of genes delivered directly into liver and muscle by lentiviral vectors. Nat Genet. 1997;17:314–317. doi: 10.1038/ng1197-314. [DOI] [PubMed] [Google Scholar]
  81. Kay MA, Glorioso JC, Naldini L. Viral vectors for gene therapy: The art of turning infectious agents into vehicles of therapeutics. Nat Med. 2001;7:33–40. doi: 10.1038/83324. [DOI] [PubMed] [Google Scholar]
  82. Kerr TP, Sewry CA, Robb SA, Roberts RG. Long mutant dystrophins and variable phenotypes: Evasion of nonsense-mediated decay? Hum Genet. 2001;109:402–407. doi: 10.1007/s004390100598. [DOI] [PubMed] [Google Scholar]
  83. Khurana TS, Prendergast RA, Alameddine HS, Tome FM, Fardeau M, Arahata K, Sugita H, Kunkel LM. Absence of extraocular muscle pathology in Duchenne’s muscular dystrophy: Role for calcium homeostasis in extraocular muscle sparing. J Exp Med. 1995;182:467–475. doi: 10.1084/jem.182.2.467. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Koenig M, Monaco AP, Kunkel LM. The complete sequence of dystrophin predicts a rod-shaped cytoskeletal protein. Cell. 1988;53:219–226. doi: 10.1016/0092-8674(88)90383-2. [DOI] [PubMed] [Google Scholar]
  85. Koenig M, Beggs AH, Moyer M, Scherpf S, Heindrich K, Bettecken T, Meng G, Muller CR, Lindlof M, Kaariainen H. The molecular basis for Duchenne versus Becker muscular dystrophy: Correlation of severity with type of deletion. Am J Hum Genet. 1989;45:498–506. [PMC free article] [PubMed] [Google Scholar]
  86. Kornegay JN, Tuler SM, Miller DM, Levesque DC. Muscular dystrophy in a litter of golden retriever dogs. Musc Nerve. 1988;11:1056–1064. doi: 10.1002/mus.880111008. [DOI] [PubMed] [Google Scholar]
  87. Kornegay JN, Sharp NJ, Bartlett RJ, Van Camp SD, Burt CT, Hung WY, Kwock L, Roses AD. Golden retriever muscular dystrophy: Monitoring for success. Adv Exp Med Biol. 1990;280:267–272. doi: 10.1007/978-1-4684-5865-7_30. [DOI] [PubMed] [Google Scholar]
  88. Kucia M, Reca R, Jala VR, Dawn B, Ratajczak J, Ratajczak MZ. Bone marrow as a home of heterogenous populations of nonhematopoietic stem cells. Leukemia. 2005;19:1118–1127. doi: 10.1038/sj.leu.2403796. [DOI] [PubMed] [Google Scholar]
  89. Kuhr CS, Lupu M, Storb R. Hematopoietic cell transplantation directly into dystrophic muscle fails to reconstitute satellite cells and myofibers. Biol Blood Marr Transpl. 2007a;13:886–888. doi: 10.1016/j.bbmt.2007.04.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Kuhr CS, Yunusov M, Sale G, Loretz C, Storb R. Long-term tolerance to kidney allografts in a preclinical canine model. Transplantation. 2007b;84:545–547. doi: 10.1097/01.tp.0000270325.84036.52. [DOI] [PubMed] [Google Scholar]
  91. LaBarge MA, Blau HM. Biological progression from adult bone marrow to mononucleate muscle stem cell to multinucleate muscle fiber in response to injury. Cell. 2002;111:589–601. doi: 10.1016/s0092-8674(02)01078-4. [DOI] [PubMed] [Google Scholar]
  92. Lakshmipathy U, Verfaillie C. Stem cell plasticity. Blood Rev. 2005;19:29–38. doi: 10.1016/j.blre.2004.03.001. [DOI] [PubMed] [Google Scholar]
  93. Lapidos KA, Kakkar R, McNally EM. The dystrophin glycoprotein complex: Signaling strength and integrity for the sarcolemma. Circ Res. 2004;94:1023–1031. doi: 10.1161/01.RES.0000126574.61061.25. [DOI] [PubMed] [Google Scholar]
  94. Lee-Pullen TF, Bennett AL, Beilharz MW, Grounds MD, Sammels LM. Superior survival and proliferation after transplantation of myoblasts obtained from adult mice compared with neonatal mice. Transplantation. 2004;78:1172–1176. doi: 10.1097/01.tp.0000137936.75203.b4. [DOI] [PubMed] [Google Scholar]
  95. Lidov HG, Selig S, Kunkel LM. Dp140: A novel 140 kDa CNS transcript from the dystrophin locus. Hum Mol Genet. 1995;4:329–335. doi: 10.1093/hmg/4.3.329. [DOI] [PubMed] [Google Scholar]
  96. Liu M, Yue Y, Harper SQ, Grange RW, Chamberlain JS, Duan D. Adeno-associated virus-mediated microdystrophin expression protects young mdx muscle from contraction-induced injury. Mol Ther. 2005;11:245–256. doi: 10.1016/j.ymthe.2004.09.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Lynch GS, Hinkle RT, Chamberlain JS, Brooks SV, Faulkner JA. Force and power output of fast and slow skeletal muscles from mdx mice 6–28 months old. J Physiol. 2001;535:591–600. doi: 10.1111/j.1469-7793.2001.00591.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. MacKenzie TC, Kobinger GP, Kootstra NA, Radu A, Sena-Esteves M, Bouchard S, Wilson JM, Verma IM, Flake AW. Efficient transduction of liver and muscle after in utero injection of lentiviral vectors with different pseudotypes. Mol Ther. 2002;6:349–358. doi: 10.1006/mthe.2002.0681. [DOI] [PubMed] [Google Scholar]
  99. Manzur AY, Kuntzer T, Pike M, Swan A. Glucocorticoid corticosteroids for Duchenne muscular dystrophy. update of Cochrane Database Syst Rev 2004. 2008;(2):CD003725. doi: 10.1002/14651858.CD003725.pub2. Cochrane Db Syst Rev, CD003725. [DOI] [PubMed] [Google Scholar]
  100. McClorey G, Moulton HM, Iversen PL, Fletcher S, Wilton SD. Antisense oligonucleotide-induced exon skipping restores dystrophin expression in vitro in a canine model of DMD. Gene Ther. 2006;13:1373–1381. doi: 10.1038/sj.gt.3302800. [DOI] [PubMed] [Google Scholar]
  101. Mendell JR, Kissel JT, Amato AA, King W, Signore L, Prior TW, Sahenk Z, Benson S, McAndrew PE, Rice R. Myoblast transfer in the treatment of Duchenne’s muscular dystrophy. New Engl J Med. 1995;333:832–838. doi: 10.1056/NEJM199509283331303. [DOI] [PubMed] [Google Scholar]
  102. Miller RG, Sharma KR, Pavlath GK, Gussoni E, Mynhier M, Lanctot AM, Greco CM, Steinman L, Blau HM. Myoblast implantation in Duchenne muscular dystrophy: The San Francisco study. Musc Nerve. 1997;20:469–478. doi: 10.1002/(sici)1097-4598(199704)20:4<469::aid-mus10>3.0.co;2-u. [DOI] [PubMed] [Google Scholar]
  103. Muntoni F, Torelli S, Ferlini A. Dystrophin and mutations: One gene, several proteins, multiple phenotypes. Lancet Neurol. 2003;2:731–740. doi: 10.1016/s1474-4422(03)00585-4. [DOI] [PubMed] [Google Scholar]
  104. Pacak CA, Mah CS, Thattaliyath BD, Conlon TJ, Lewis MA, Cloutier DE, Zolotukhin I, Tarantal AF, Byrne BJ. Recombinant adeno-associated virus serotype 9 leads to preferential cardiac transduction in vivo. Circ Res. 2006;99:e3–e9. doi: 10.1161/01.RES.0000237661.18885.f6. [DOI] [PubMed] [Google Scholar]
  105. Parker MH, Kuhr C, Tapscott SJ, Storb R. Hematopoietic cell transplantation provides an immune-tolerant platform for myoblast transplantation in dystrophic dogs. Mol Ther. 2008;16:1340–1346. doi: 10.1038/mt.2008.102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Partridge T. The current status of myoblast transfer. Neurol Sci. 2000;21 (Suppl):S939–S942. doi: 10.1007/s100720070007. [DOI] [PubMed] [Google Scholar]
  107. Partridge T, Lu QL, Morris G, Hoffman E. Is myoblast transplantation effective? Nat Med. 1998;4:1208–1209. doi: 10.1038/3167. [DOI] [PubMed] [Google Scholar]
  108. Phelps SF, Hauser MA, Cole NM, Rafael JA, Hinkle RT, Faulkner JA, Chamberlain JS. Expression of full-length and truncated dystrophin mini-genes in transgenic mdx mice. Hum Mol Genet. 1995;4:1251–1258. doi: 10.1093/hmg/4.8.1251. [DOI] [PubMed] [Google Scholar]
  109. Pirenne J, Kawai M. Tolerogenic protocols for intestinal transplantation. Transpl Immunol. 2004;13:131–137. doi: 10.1016/j.trim.2004.05.005. [DOI] [PubMed] [Google Scholar]
  110. Porter JD. Commentary: Extraocular muscle sparing in muscular dystrophy: A critical evaluation of potential protective mechanisms. Neuromusc Dis. 1998;8:198–203. doi: 10.1016/s0960-8966(98)00015-7. [DOI] [PubMed] [Google Scholar]
  111. Porter JD, Merriam AP, Leahy P, Gong B, Khanna S. Dissection of temporal gene expression signatures of affected and spared muscle groups in dystrophin-deficient (mdx) mice. Hum Mol Genet. 2003;12:1813–1821. doi: 10.1093/hmg/ddg197. [DOI] [PubMed] [Google Scholar]
  112. Qiao C, Li J, Jiang J, Zhu X, Wang B, Li J, Xiao X. Myostatin propeptide gene delivery by adeno-associated virus serotype 8 vectors enhances muscle growth and ameliorates dystrophic phenotypes in mdx mice. Hum Gene Ther. 2008a;19:241–254. doi: 10.1089/hum.2007.159. [DOI] [PubMed] [Google Scholar]
  113. Qiao C, Li J, Zheng H, Bogan J, Li J, Yuan Z, Zhang C, Bogan D, Kornegay J, Xiao X. Hydrodynamic limb vein injection of AAV8 canine myostatin propeptide gene in normal dogs enhances muscle growth. Hum Gene Ther 2008. 2008b Oct 1; doi: 10.1089/hum.2008.135. [Epub ahead of print] [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Rando TA. The dystrophin-glycoprotein complex, cellular signaling, and the regulation of cell survival in the muscular dystrophies. Musc Nerve. 2001;24:1575–1594. doi: 10.1002/mus.1192. [DOI] [PubMed] [Google Scholar]
  115. Rando TA. Non-viral gene therapy for Duchenne muscular dystrophy: Progress and challenges. Biochim Biophys Acta. 2007;1772:263–271. doi: 10.1016/j.bbadis.2006.07.009. [DOI] [PubMed] [Google Scholar]
  116. Rodino-Klapac LR, Chicoine LG, Kaspar BK, Mendell JR. Gene therapy for Duchenne muscular dystrophy: Expectations and challenges. Arch Neurol. 2007;64:1236–1241. doi: 10.1001/archneur.64.9.1236. [DOI] [PubMed] [Google Scholar]
  117. Romero NB, Braun S, Benveniste O, Leturcq F, Hogrel JY, Morris GE, Barois A, Eymard B, Payan C, Ortega V, Boch AL, Lejean L, Thioudellet C, Mourot B, Escot C, Choquel A, Recan D, Kaplan JC, Dickson G, Klatzmann D, Molinier-Frenckel V, Guillet JG, Squiban P, Herson S, Fardeau M. Phase I study of dystrophin plasmid-based gene therapy in Duchenne/Becker muscular dystrophy. Hum Gene Ther. 2004;15:1065–1076. doi: 10.1089/hum.2004.15.1065. [DOI] [PubMed] [Google Scholar]
  118. Rutledge EA, Halbert CL, Russell DW. Infectious clones and vectors derived from adeno-associated virus (AAV) serotypes other than AAV type 2. J Virol. 1998;72:309–319. doi: 10.1128/jvi.72.1.309-319.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  119. Sampaolesi M, Torrente Y, Innocenzi A, Tonlorenzi R, D’Antona G, Pellegrino MA, Barresi R, Bresolin N, De Angelis MG, Campbell KP, Bottinelli R, Cossu G. Cell therapy of alpha-sarcoglycan null dystrophic mice through intra-arterial delivery of mesoangioblasts. Science. 2003;301:487–492. doi: 10.1126/science.1082254. [DOI] [PubMed] [Google Scholar]
  120. Sampaolesi M, Blot S, D’Antona G, Granger N, Tonlorenzi R, Innocenzi A, Mognol P, Thibaud JL, Galvez BG, Barthelemy I, Perani L, Mantero S, Guttinger M, Pansarasa O, Rinaldi C, Cusella De Angelis MG, Torrente Y, Bordignon C, Bottinelli R, Cossu G. Mesoangioblast stem cells ameliorate muscle function in dystrophic dogs. Nature. 2006;444:574–579. doi: 10.1038/nature05282. [DOI] [PubMed] [Google Scholar]
  121. Schatzberg SJ, Olby NJ, Breen M, Anderson LV, Langford CF, Dickens HF, Wilton SD, Zeiss CJ, Binns MM, Kornegay JN, Morris GE, Sharp NJ. Molecular analysis of a spontaneous dystrophin ‘knockout’ dog. Neuromusc Dis. 1999;9:289–295. doi: 10.1016/s0960-8966(99)00011-5. [DOI] [PubMed] [Google Scholar]
  122. Schiedner G, Morral N, Parks RJ, Wu Y, Koopmans SC, Langston C, Graham FL, Beaudet AL, Kochanek S. Genomic DNA transfer with a high-capacity adenovirus vector results in improved in vivo gene expression and decreased toxicity [erratum in Nat Genet 18:298] Nat Genet. 1998;18:180–183. doi: 10.1038/ng0298-180. [DOI] [PubMed] [Google Scholar]
  123. Scime A, Rudnicki MA. Molecular-targeted therapy for Duchenne muscular dystrophy: Progress and potential. Mol Diagn Ther. 2008;12:99–108. doi: 10.1007/BF03256275. [DOI] [PubMed] [Google Scholar]
  124. Scott JM, Li S, Harper SQ, Welikson R, Bourque D, DelloRusso C, Hauschka SD, Chamberlain JS. Viral vectors for gene transfer of micro-, mini-, or full-length dystrophin. Neuromusc Dis. 2002;12 (Suppl 1):S23–S29. doi: 10.1016/s0960-8966(02)00078-0. [DOI] [PubMed] [Google Scholar]
  125. Sharp NJ, Kornegay JN, Van Camp SD, Herbstreith MH, Secore SL, Kettle S, Hung WY, Constantinou CD, Dykstra MJ, Roses AD, Bartlett RJ. An error in dystrophin mRNA processing in golden retriever muscular dystrophy, an animal homologue of Duchenne muscular dystrophy. Genomics. 1992;13:115–121. doi: 10.1016/0888-7543(92)90210-j. [DOI] [PubMed] [Google Scholar]
  126. Shimatsu Y, Katagiri K, Furuta T, Nakura M, Tanioka Y, Yuasa K, Tomohiro M, Kornegay JN, Nonaka I, Takeda S. Canine X-linked muscular dystrophy in Japan (CXMDJ) Exp Anim. 2003;52:93–97. doi: 10.1538/expanim.52.93. [DOI] [PubMed] [Google Scholar]
  127. Sicinski P, Geng Y, Ryder-Cook AS, Barnard EA, Darlison MG, Barnard PJ. The molecular basis of muscular dystrophy in the mdx mouse: A point mutation. Science. 1989;244:1578–1580. doi: 10.1126/science.2662404. [DOI] [PubMed] [Google Scholar]
  128. Sinn PL, Sauter SL, McCray PB., Jr Gene therapy progress and prospects: Development of improved lentiviral and retroviral vectors design, biosafety, and production. Gene Ther. 2005;12:1089–1098. doi: 10.1038/sj.gt.3302570. [DOI] [PubMed] [Google Scholar]
  129. Skuk D, Roy B, Goulet M, Chapdelaine P, Bouchard JP, Roy R, Dugre FJ, Lachance JG, Deschenes L, Helene S, Sylvain M, Tremblay JP. Dystrophin expression in myofibers of Duchenne muscular dystrophy patients following intramuscular injections of normal myogenic cells. Mol Ther. 2004;9:475–482. doi: 10.1016/j.ymthe.2003.11.023. [DOI] [PubMed] [Google Scholar]
  130. Skuk D, Goulet M, Roy B, Chapdelaine P, Bouchard JP, Roy R, Dugre FJ, Sylvain M, Lachance JG, Deschenes L, Senay H, Tremblay JP. Dystrophin expression in muscles of Duchenne muscular dystrophy patients after high-density injections of normal myogenic cells. J Neuropath Exp Neurol. 2006;65:371–386. doi: 10.1097/01.jnen.0000218443.45782.81. [DOI] [PubMed] [Google Scholar]
  131. Spencer MJ, Mellgren RL. Overexpression of a calpastatin transgene in mdx muscle reduces dystrophic pathology. Hum Mol Genet. 2002;11:2645–2655. doi: 10.1093/hmg/11.21.2645. [DOI] [PubMed] [Google Scholar]
  132. Stedman HH, Sweeney HL, Shrager JB, Maguire HC, Panettieri RA, Petrof B, Narusawa M, Leferovich JM, Sladky JT, Kelly AM. The mdx mouse diaphragm reproduces the degenerative changes of Duchenne muscular dystrophy. Nature. 1991;352:536–539. doi: 10.1038/352536a0. [DOI] [PubMed] [Google Scholar]
  133. Storb R. Allogeneic hematopoietic stem cell transplantation: Yesterday, today, and tomorrow. Exp Hemat. 2003;31:1–10. doi: 10.1016/s0301-472x(02)01020-2. [DOI] [PubMed] [Google Scholar]
  134. Storb R, Yu C, Wagner JL, Deeg HJ, Nash RA, Kiem H-P, Leisenring W, Shulman H. Stable mixed hematopoietic chimerism in DLA-identical littermate dogs given sublethal total body irradiation before and pharmacological immunosuppression after marrow transplantation. Blood. 1997;89:3048–3054. [PubMed] [Google Scholar]
  135. Su LT, Gopal K, Wang Z, Yin X, Nelson A, Kozyak BW, Burkman JM, Mitchell MA, Low DW, Bridges CR, Stedman HH. Uniform scale-independent gene transfer to striated muscle after transvenular extravasation of vector. Circulation. 2005;112:1780–1788. doi: 10.1161/CIRCULATIONAHA.105.534008. [DOI] [PubMed] [Google Scholar]
  136. Thomas GD, Sander M, Lau KS, Huang PL, Stull JT, Victor RG. Impaired metabolic modulation of alpha-adrenergic vasoconstriction in dystrophin-deficient skeletal muscle. Proc Natl Acad Sci U S A. 1998;95:15090–15095. doi: 10.1073/pnas.95.25.15090. [DOI] [PMC free article] [PubMed] [Google Scholar]
  137. Tidball JG, Wehling-Henricks M. Evolving therapeutic strategies for Duchenne muscular dystrophy: Targeting downstream events. Ped Res. 2004;56:831–841. doi: 10.1203/01.PDR.0000145578.01985.D0. [DOI] [PubMed] [Google Scholar]
  138. Tidball JG, Wehling-Henricks M. Damage and inflammation in muscular dystrophy: Potential implications and relationships with autoimmune myositis. Curr Opin Rheum. 2005;17:707–713. doi: 10.1097/01.bor.0000179948.65895.1a. [DOI] [PubMed] [Google Scholar]
  139. Torrente Y, Belicchi M, Sampaolesi M, Pisati F, Meregalli M, D’Antona G, Tonlorenzi R, Porretti L, Gavina M, Mamchaoui K, Pellegrino MA, Furling D, Mouly V, Butler-Browne GS, Bottinelli R, Cossu G, Bresolin N. Human circulating AC133(+) stem cells restore dystrophin expression and ameliorate function in dystrophic skeletal muscle. J Clin Invest. 2004;114:182–195. doi: 10.1172/JCI20325. [DOI] [PMC free article] [PubMed] [Google Scholar]
  140. Tremblay JP, Malouin F, Roy R, Huard J, Bouchard JP, Satoh A, Richards CL. Results of a triple blind clinical study of myoblast transplantations without immunosuppressive treatment in young boys with Duchenne muscular dystrophy. Cell Transpl. 1993;2:99–112. doi: 10.1177/096368979300200203. [DOI] [PubMed] [Google Scholar]
  141. Tyler KL. Origins and early descriptions of “Duchenne muscular dystrophy. Musc Nerve. 2003;28:402–422. doi: 10.1002/mus.10435. [DOI] [PubMed] [Google Scholar]
  142. Valentine BA, Cooper BJ. Canine X-linked muscular dystrophy: Selective involvement of muscles in neonatal dogs. Neuromusc Dis. 1991;1:31–38. doi: 10.1016/0960-8966(91)90040-y. [DOI] [PubMed] [Google Scholar]
  143. Valentine BA, Cooper BJ, Cummings JF, de Lahunta A. Progressive muscular dystrophy in a golden retriever dog: Light microscope and ultrastructural features at 4 and 8 months. Acta Neuropathol. 1986;71:301–310. doi: 10.1007/BF00688053. [DOI] [PubMed] [Google Scholar]
  144. Valentine BA, Cooper BJ, de Lahunta A, O’Quinn R, Blue JT. Canine X-linked muscular dystrophy: An animal model of Duchenne muscular dystrophy (Clinical studies) J Neurol Sci. 1988;88:69–81. doi: 10.1016/0022-510x(88)90206-7. [DOI] [PubMed] [Google Scholar]
  145. Valentine BA, Winand NJ, Pradhan D, Moise NS, de Lahunta A, Kornegay JN, Cooper BJ. Canine X-linked muscular dystrophy as an animal model of Duchenne muscular dystrophy: A review. Am J Med Genet. 1992;42:352–356. doi: 10.1002/ajmg.1320420320. [DOI] [PubMed] [Google Scholar]
  146. van Deutekom JC, van Ommen GJ. Advances in Duchenne muscular dystrophy gene therapy. Nat Rev Genet. 2003;4:774–783. doi: 10.1038/nrg1180. [DOI] [PubMed] [Google Scholar]
  147. van Deutekom JC, Janson AA, Ginjaar IB, Frankhuizen WS, Aartsma-Rus A, Bremmer-Bout M, Den Dunnen JT, Koop K, van der Kooi AJ, Goemans NM, de Kimpe SJ, Ekhart PF, Venneker EH, Platenburg GJ, Verschuuren JJ, van Ommen GJ. Local dystrophin restoration with antisense oligonucleotide PRO051. New Engl J Med. 2007;357:2677–2686. doi: 10.1056/NEJMoa073108. [DOI] [PubMed] [Google Scholar]
  148. Vitiello L, Bassi N, Campagnolo P, Zaccariotto E, Occhi G, Malerba A, Pigozzo S, Reggiani C, Ausoni S, Zaglia T, Gamba P, Baroni MD, Ditadi AP. In vivo delivery of naked antisense oligos in aged mdx mice: Analysis of dystrophin restoration in skeletal and cardiac muscle. Neuromusc Dis. 2008;18:597–605. doi: 10.1016/j.nmd.2008.05.011. [DOI] [PubMed] [Google Scholar]
  149. Wagner KR, Lechtzin N, Judge DP. Current treatment of adult Duchenne muscular dystrophy. Biochim Biophys Acta. 2007;1772:229–237. doi: 10.1016/j.bbadis.2006.06.009. [DOI] [PubMed] [Google Scholar]
  150. Wakefield PM, Tinsley JM, Wood MJ, Gilbert R, Karpati G, Davies KE. Prevention of the dystrophic phenotype in dystrophin/utrophin-deficient muscle following adenovirus-mediated transfer of a utrophin minigene. Gene Ther. 2000;7:201–204. doi: 10.1038/sj.gt.3301066. [DOI] [PubMed] [Google Scholar]
  151. 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 Biotech. 2005;23:321–328. doi: 10.1038/nbt1073. [DOI] [PubMed] [Google Scholar]
  152. Wang Z, Zhu T, Rehman KK, Bertera S, Zhang J, Chen C, Papworth G, Watkins S, Trucco M, Robbins PD, Li J, Xiao X. Widespread and stable pancreatic gene transfer by adeno-associated virus vectors via different routes. Diabetes. 2006;55:875–884. doi: 10.2337/diabetes.55.04.06.db05-0927. [DOI] [PubMed] [Google Scholar]
  153. Wang Z, Allen JM, Riddell SR, Gregorevic P, Storb R, Tapscott SJ, Chamberlain JS, Kuhr CS. Immunity to adeno-associated virus-mediated gene transfer in a random-bred canine model of Duchenne muscular dystrophy. Hum Gene Ther. 2007a;18:18–26. doi: 10.1089/hum.2006.093. [DOI] [PubMed] [Google Scholar]
  154. Wang Z, Kuhr CS, Allen JM, Blankinship M, Gregorevic P, Chamberlain JS, Tapscott SJ, Storb R. Sustained AAV-mediated dystrophin expression in a canine model of Duchenne muscular dystrophy with a brief course of immunosuppression. Mol Ther. 2007b;15:1160–1166. doi: 10.1038/sj.mt.6300161. [DOI] [PubMed] [Google Scholar]
  155. Warrington KH, Jr, Herzog RW. Treatment of human disease by adeno-associated viral gene transfer. Hum Genet. 2006;119:571–603. doi: 10.1007/s00439-006-0165-6. [DOI] [PubMed] [Google Scholar]
  156. Wehling-Henricks M, Lee JJ, Tidball JG. Prednisolone decreases cellular adhesion molecules required for inflammatory cell infiltration in dystrophin-deficient skeletal muscle. Neuromusc Dis. 2004;14:483–490. doi: 10.1016/j.nmd.2004.04.008. [DOI] [PubMed] [Google Scholar]
  157. Welch EM, Barton ER, Zhuo J, Tomizawa Y, Friesen WJ, Trifillis P, Paushkin S, Patel M, Trotta CR, Hwang S, Wilde RG, Karp G, Takasugi J, Chen G, Jones S, Ren H, Moon YC, Corson D, Turpoff AA, Campbell JA, Conn MM, Khan A, Almstead NG, Hedrick J, Mollin A, Risher N, Weetall M, Yeh S, Branstrom AA, Colacino JM, Babiak J, Ju WD, Hirawat S, Northcutt VJ, Miller LL, Spatrick P, He F, Kawana M, Feng H, Jacobson A, Peltz SW, Sweeney HL. PTC124 targets genetic disorders caused by nonsense mutations. Nature. 2007;447:87–91. doi: 10.1038/nature05756. [DOI] [PubMed] [Google Scholar]
  158. Wells DJ, Wells KE, Asante EA, Turner G, Sunada Y, Campbell KP, Walsh FS, Dickson G. Expression of human full-length and minidystrophin in transgenic mdx mice: Implications for gene therapy of Duchenne muscular dystrophy. Hum Mol Genet. 1995;4:1245–1250. doi: 10.1093/hmg/4.8.1245. [DOI] [PubMed] [Google Scholar]
  159. Wiendl H, Hohlfeld R, Kieseier BC. Immunobiology of muscle: Advances in understanding an immunological microenvironment. Trends Immunol. 2005;26:373–380. doi: 10.1016/j.it.2005.05.003. [DOI] [PubMed] [Google Scholar]
  160. Wilton SD, Lloyd F, Carville K, Fletcher S, Honeyman K, Agrawal S, Kole R. Specific removal of the nonsense mutation from the mdx dystrophin mRNA using antisense oligonucleotides. Neuromusc Dis. 1999;9:330–338. doi: 10.1016/s0960-8966(99)00010-3. [DOI] [PubMed] [Google Scholar]
  161. Wolff J, Lewis DL, Herweijer H, Hegge J, Hagstrom J. Non-viral approaches for gene transfer. Acta Myol. 2005;24:202–208. [PubMed] [Google Scholar]
  162. Yue Y, Ghosh A, Long C, Bostick B, Smith BF, Kornegay JN, Duan D. A single intravenous injection of adeno-associated virus serotype-9 leads to whole body skeletal muscle transduction in dogs. Mol Ther. 2008;16:1944–1952. doi: 10.1038/mt.2008.207. [DOI] [PMC free article] [PubMed] [Google Scholar]
  163. Zhao J, Bauman WA, Huang R, Caplan AJ, Cardozo C. Oxandrolone blocks glucocorticoid signaling in an androgen receptor-dependent manner. Steroids. 2004;69:357–366. doi: 10.1016/j.steroids.2004.01.006. [DOI] [PubMed] [Google Scholar]

RESOURCES