Sei sulla pagina 1di 14

Human Molecular Genetics, 2011, Vol.

20, Review Issue 1 R28–R41


doi:10.1093/hmg/ddr110
Advance Access published on March 23, 2011

Genetic therapy for the nervous system


William J. Bowers 1, Xandra O. Breakefield 2,3,∗ and Miguel Sena-Esteves 4
1
Department of Neurology, Center for Neural Development and Disease, University of Rochester,
School of Medicine and Dentistry, Rochester, NY 14642, USA, 2Neuroscience Center and Molecular Neurogenetics
Unit, Department of Neurology and 3Center for Molecular Imaging Research, Department of Radiology,
Massachusetts General Hospital and Program in Neuroscience, Harvard Medical School, Boston, MA 02114, USA
and 4Department of Neurology, Gene Therapy Center, Interdisciplinary Graduate Program, University of

Downloaded from https://academic.oup.com/hmg/article-abstract/20/R1/R28/631150 by guest on 15 March 2019


Massachusetts Medical School, Worcester, MA 01605, USA

Received February 1, 2011; Revised and Accepted March 11, 2011

Genetic therapy is undergoing a renaissance with expansion of viral and synthetic vectors, use of oligonu-
cleotides (RNA and DNA) and sequence-targeted regulatory molecules, as well as genetically modified
cells, including induced pluripotent stem cells from the patients themselves. Several clinical trials for neuro-
logic syndromes appear quite promising. This review covers genetic strategies to ameliorate neurologic
syndromes of different etiologies, including lysosomal storage diseases, Alzheimer’s disease and other amy-
loidopathies, Parkinson’s disease, spinal muscular atrophy, amyotrophic lateral sclerosis and brain tumors.
This field has been propelled by genetic technologies, including identifying disease genes and disruptive
mutations, design of genomic interacting elements to regulate transcription and splicing of specific precur-
sor mRNAs and use of novel non-coding regulatory RNAs. These versatile new tools for manipulation of
genetic elements provide the ability to tailor the mode of genetic intervention to specific aspects of a disease
state.

INTRODUCTION METHODS OF GENETIC INTERVENTION


Genetic therapy covers a range of methods for modifying the Routes of access
nervous system, including delivery of genes, sequence-
targeted regulatory molecules, genetically modified cells and Entrance to the central nervous system (CNS) is limited by the
oligonucleotides. In this review, we focus on ways to change skeletal structures that enclose it and the blood – brain barrier
the genetic status of the nervous system, including routes of (BBB), which plays a central role in regulating the chemical
access and modes of delivery. Examples are provided of strat- microenvironment of the CNS by preventing simple diffusion
egies used for a number of diseases, including recessive loss of of most small molecules and proteins from the bloodstream.
function conditions [lysosomal storage diseases and spinal The physical barrier properties of the BBB are a result of
muscular atrophy (SMA)], dominant toxic mutations [amyo- the tight junctions between brain microvascular endothelial
trophic lateral sclerosis (ALS)] and conditions of mixed etiol- cells and low vesicular transport (1; Fig. 2). As a result,
ogy [Alzheimer’s disease (AD), Parkinson’s disease (PD) and most molecular traffic in and out of the CNS is tightly regu-
brain tumors] (Fig. 1). Neurodegenerative diseases caused by lated by specialized transporter systems present on the
triplet nucleotide repeats are discussed in an article on luminal and abluminal membranes of the endothelial cells,
siRNAs in this issue. Many of these strategies have proven with the exception of small solutes such O2 and CO2 gases
very promising in preclinical studies in mouse and larger and some lipophilic molecules (e.g. ethanol) that can diffuse
animal models, and a substantial number are now being eval- freely across cellular membranes. Other components of the
uated in clinical trials. BBB are the basal lamina, astrocyte end-feet that surround


To whom correspondence should be addressed at: Molecular Neurogenetics Unit, Massachusetts General Hospital-East, 13th Street, Building 149,
Charlestown, MA 02129, USA. Tel: +1 6177265728; Fax: +1 6177241537; Email: breakefield@hms.harvard.edu

# The Author 2011. Published by Oxford University Press. All rights reserved.
For Permissions, please email: journals.permissions@oup.com
Human Molecular Genetics, 2011, Vol. 20, Review Issue 1 R29

Downloaded from https://academic.oup.com/hmg/article-abstract/20/R1/R28/631150 by guest on 15 March 2019


Figure 1. Therapeutic strategies based on genetic etiology. (A) Recessive disease. In the case of a recessively inherited disease where both alleles (or one on the
X chromosome in males) of a gene are mutated, the goal is to replace the defective gene with a functional counterpart or to correct the defect. Typically, a vector
is used to deliver a promoter–cDNA expression cassette which either integrates into the genome (1), e.g. lentivirus vector, or resides as an extrachromosomal
element (4), e.g. AAV vector. Other approaches include attempts to achieve homologous recombination with the resident allele (4) or, if appropriate, to correct
the transcript through transplicing of the precursor mRNA (3). (B) Dominant disease. In the case of dominantly inherited diseases with only one defective neu-
rotoxic gene, a common strategy is to block expression of the mutant message through RNAi (1) which can be delivered either as an shRNA encoded in a vector
or as an oligonucleotide. Another approach is to try to neutralize damage caused by the mutant protein through the use of targeted antibodies (2). (C) Multi-
factorial dysfunction. For many neurodegenerative diseases, there is no one gene that can be targeted to alleviate the condition. Approaches include trying
to supplement the neurotransmitter capacity of ‘sick’ neurons by providing them with enzymes to generate more of their normal neurotransmitter for release
at synapses (1). Alternately, the neurotransmitter profile of interacting neurons can be altered to change their output from excitatory to inhibitory, or vice
versa (2). In another common scenario, cells in the vicinity are genetically modified to produce a growth factor which is supportive for the sick neurons (3).

all CNS blood vessels and pericytes. Interestingly, pericytes widespread transduction in the neonatal mouse brain (5,7),
appear to play a central role in regulating BBB permeability but the distribution pattern is more restricted in adult
via their influence over endothelial cells and astrocytes (2). animals (6,8,9). This suggests that there may be additional bar-
Not surprisingly, the BBB has proved exceptionally efficient riers to distribution of gene transfer vectors from CSF into
in excluding the vast majority of gene transfer vehicles adult brains, or that commonly used injection procedures
from reaching the CNS via the vasculature. As a result, most disrupt normal CSF flow preventing widespread distribution.
CNS gene therapy approaches have utilized direct infusion Delivery of recombinant lysosomal enzymes into CSF
of gene transfer vectors into the brain parenchyma to appears to be effective in providing therapeutic levels of
target disease-relevant structures. The distribution of viral these enzymes throughout the CNS (10,11). Similarly, CSF
vectors in the brain can be improved considerably by infusion of antisense oligonucleotides (ASOs) directed at
convection-enhanced delivery (CED), a slow pressurized infu- altering the splicing pattern of the SMN2 gene has been dra-
sion of a larger volume (3,4). Nonetheless, transduction of matically effective in a mouse model of SMA (see below).
CNS cells after intraparenchymal injection of viral vectors An ideal route of entry into the CNS to achieve widespread
remains mostly limited to the targeted structure. gene transfer would be through the vasculature. Until recently,
An alternative route of entry into the CNS is by delivery of the only exception to the BBB-imposed block to gene therapy
the gene transfer vectors directly into cerebrospinal fluid vectors was PEGylated immunoliposomes (PILs) formulated
(CSF) via either the lateral ventricles (5) or intrathecal space with monoclonal antibodies specific for receptors, such as
(6). This approach is exceptionally effective in achieving transferrin and insulin receptors which mediate transcytosis
R30 Human Molecular Genetics, 2011, Vol. 20, Review Issue 1

Downloaded from https://academic.oup.com/hmg/article-abstract/20/R1/R28/631150 by guest on 15 March 2019


Figure 2. Routes of gene delivery to the CNS. CED of viral vectors into the brain improves considerably their distribution in target structures and hence transduc-
tion volumes. This technique can yield volumes of transduced cell distribution 3 –3.5-fold larger than the infused volume, which is highly significant for human
applications. Viral vectors or secreted transgene products (growth factors, lysosomal enzymes) can be further distributed from the primary target structure by
axonal transport (top left diagram). Infusion of recombinant proteins or oligonucleotides into the brain ventricular system, or intrathecal space, leads to wide-
spread CNS distribution via CSF flow. An alternative strategy is to use viral vectors to engineer ependymal cells lining the ventricles or choroid plexus cells to
secrete therapeutic proteins into CSF (top right diagram). The BBB with its many constituents has thwarted most gene transfer vehicles from entering the brain
from the vasculature. In recent years, PILs and a new generation of viral vectors (AAV and SV40) have been shown to mediate efficient CNS gene transfer after
i.v. infusion in newborn and adult animals (bottom right diagram).

of their ligands across the BBB. These PILs appear to be quite factors, cytokines, tumor killing agents) are to target gene
effective in delivering expression plasmids (with expression transfer to brain microcapillary endothelial cells (17), or use
under cell type-specific promoters) and RNAi to normal ex vivo genetically modified stem cells [hematopoietic stem
brain or brain tumors (12). In a parallel approach, therapeutic cells (HSCs), neural stem cells, mesenchymal stem cells]
proteins are delivered to the CNS using chimeric recombinant which will themselves, or their progeny in the case of HSCs,
molecules, e.g. growth factors, single-chain antibodies or lyso- migrate within the brain to regions of injury or tumors
somal enzymes, fused to receptor-targeting monoclonal anti- (18 –20). In fact, the whole nervous system can be transduced
bodies (12) or ligands, such as transferrin (13). Recently, with a gene for the lifespan by injecting a lentivirus vector into
adeno-associated virus 9 (AAV9) vectors have been found to the amniotic sac of mouse embryos before the neural groove
enter the CNS of neonatal mice and young cats after intravas- has closed (21).
cular (i.v.) infusion and to transduce large numbers of glia and
motor neurons in the spinal cord (14,15). Transduction of
other neuronal populations in the brain is found mostly in DNA/RNA
the hippocampus and Purkinje cells in the cerebellum (14). The use of non-viral nucleic acid delivery to selectively
SV40 recombinant vectors also appear to mediate efficient modify cellular processes within the brain presents a number
gene transfer to certain regions of the CNS after i.v. infusion of challenges, including target cell specificity and transient
in adult mice combined with intraperitoneal mannitol infusion gene expression duration. Pardridge and colleagues (22)
(16). Alternative approaches to deliver secretable therapeutic have developed PILs target to the brain when administered
proteins to the CNS (e.g. lysosomal enzymes, growth intravenously (see above). Stachowiak et al. (23) have recently
Human Molecular Genetics, 2011, Vol. 20, Review Issue 1 R31

reported the use of organically modified silica-based nanopar- EXAMPLES OF APPLICATIONS TO SPECIFIC
ticles to induce neurogenesis within the subventricular zone of DISEASES
adult mice via intraventricular delivery of DNA encoding a
recombinant nuclear form of fibroblast growth factor Lysosomal storage diseases
receptor-1. Electroporation-based nucleic acid transfection
has also been extensively documented in the prenatal and post- Lysosomal storage diseases are typically, but not exclu-
natal rodent brain [reviewed by De Vry (24)], providing the sively, childhood diseases resulting from a genetic deficiency
means to genetically modify significant numbers of neurons in a lysosomal enzyme involved in a particular metabolic
(25,26) within the living brain (27). pathway that results in lysosomal accumulation of its sub-
strate(s). Many of these enzymes can be secreted from
cells genetically engineered to overexpress them and then
taken up by enzyme-deficient cells and correctly targeted

Downloaded from https://academic.oup.com/hmg/article-abstract/20/R1/R28/631150 by guest on 15 March 2019


Viral vectors to lysosomes via mannose-6-phosphate receptors where
Recombinant viral vector systems remain the most efficient they degrade the stored substrate(s). This mechanism,
vehicles to achieve long-term stable gene expression in the known as cross-correction (38), is the basis for enzyme
CNS. Over the years, many different viral vector systems replacement therapy which is now available for a subset of
have been investigated for this purpose, including those these diseases without neurological involvement. Unfortu-
derived from herpes simplex virus type 1 (HSV-1), adeno- nately, the BBB prevents recombinant lysosomal enzymes
virus, AAV, lentiviruses—such as HIV-1, feline immunode- infused peripherally from entering the CNS. Alternative
ficiency virus or equine infectious anemia virus, and more routes of delivery and molecules are being actively investi-
recently SV40. AAV and lentivirus vectors have emerged gated (see above). These diseases are particularly well
as the vectors of choice for gene transfer to the CNS for suited for gene therapy as they are monogenic diseases
non-oncological applications as they mediate efficient long- with very well established genotype – phenotype correlations,
term gene expression with no apparent toxicity. A recent and the cross-correction mechanism makes it possible to
study has shown that AAV-mediated transgene expression devise gene delivery strategies to supply essentially the
in the primate brain continues for at least 8 years with no entire CNS with therapeutic levels of these enzymes. Two
evidence of neuroinflammation or reactive gliosis (28). main approaches have shown dramatic therapeutic effects
Moreover, several clinical trials have shown that direct infu- in animal models of lysosomal storage diseases, namely
sion of AAV2 vectors into brain parenchyma in humans is intraparenchymal infusion of recombinant viral vectors
well tolerated (29– 33). The safety profile of direct infusion (39– 46), and bone marrow transplantation with ex vivo
of lentivirus vectors into human brain remains to be evalu- lentivirus vector modified-autologous HSCs (see article by
ated. High-capacity adenovirus vectors are attractive Biffi, Cartier and Aubourg in this issue).
because of their large transgene capacity (30 kb) and The ability of focal viral vector-mediated gene delivery to
ability to mediate long-term gene expression without supply the CNS with therapeutic levels of lysosomal
immunological complications (34). These and the large enzymes is dependent on widespread distribution based on dif-
capacity HSV amplicon vectors (150 kb; 35) may be ideal fusion (47), axonal transport over long distances (48,49) and
choices to transfer large genomic regions necessary to CSF flow in the perivascular space (8). AAV-mediated
achieve physiological regulation of gene expression for par- genetic modification of highly interconnected structures in
ticular genes/sets of genes in specific cell populations in the the brain, such as deep cerebellar nuclei (50), ventral tegmen-
CNS. Full-length gene copies with intact regulatory elements tal area (51) or thalamus (4,52), leads to widespread distri-
(.100 kb) can by delivered in HSV-1 amplicon vectors bution of these enzymes in the CNS. Alternative targets,
(36), verging on the capacity to deliver virtual mini- such as the external capsule (53) or lateral ventricles (8),
chromosomes (37). The main difficulty for high-capacity take advantage of either interstitial fluid flow or CSF flow
adenovirus and HSV-1 amplicon vectors lies in difficulties for distribution of lysosomal enzymes throughout the CNS.
in large-scale production. Recombinant SV40 vectors Successful translation of AAV-based (or lentivirus-based)
display some promising properties, namely their apparent approaches to humans will likely require targeting one or
ability to cross the adult mouse BBB after i.v. infusion more of these structures to achieve therapeutic levels of
and transduce neurons and astrocytes in specific brain these enzymes throughout the CNS. Pre-clinical studies in
regions and spinal cord (16). Moreover, these vectors can large animal models of some of these diseases are highly
be administered repeatedly as they do not seem to elicit neu- encouraging (54,55). Bone marrow transplantation with
tralizing antibodies in rodents. Whether this property is lentivirus-modified autologous HSCs has shown exceptional
shared in other mammalian species remains to be deter- results in different mouse models of lysosomal storage dis-
mined. In the neuro-oncology field, recombinant vectors eases resulting in correction of pathologic findings throughout
derived from HSV-1, adenovirus, measles virus and Newcas- the CNS (18,56,57), as well as peripheral organs (58). This
tle virus appear to be the most promising in pre-clinical approach relies on genetically modified HSC-derived cells
models of brain tumors (see below). After initial attempts (macrophages in the case of CNS) trafficking to the sites of
to broadly use each viral vector system for any/all CNS disease and becoming an in situ source of recombinant
gene transfer applications, now the particular strengths of enzyme. Both of these gene therapy approaches are now
each system are being exploited to meet the specific needs being tested in human clinical trials for different lysosomal
of different applications/diseases. storage diseases.
R32 Human Molecular Genetics, 2011, Vol. 20, Review Issue 1

Alzheimer’s disease and other amyloidopathies the substantia nigra innervating the striatum in the basal
ganglia (85). Although environmental toxins and aging are
AD is a neurodegenerative disorder characterized by severe
risk factors, genetic susceptibility is critical with 16 gene
memory loss and cognitive impairment with no available
loci implicated in 5% of cases, and multifactorial hereditary
cure. Neuropathological correlates include extracellular
risk in many others (85,86). These risk factors point to oxi-
amyloid-beta peptide deposition, intracellular neurofibrillary
dative stress, mitochondrial dysfunction and reduced ability
tangle formation, decreased synaptic integrity and neuronal
to degrade abnormal proteins as etiologic factors. Although
loss. The basal forebrain cholinergic complex is significantly
L-dopa has been used for decades to relieve motor symptoms
affected by AD-related neurodegeneration (59– 63). To
of PD, it does not prevent degeneration and eventually ceases
augment cholinergic function, Tuszynski and colleagues
to be effective.
(64– 68) delivered nerve growth factor (NGF), the prototypical
New gene/cell therapy strategies have been explored exper-
neurotrophin with demonstrated neuroprotective properties,
imentally with some translated into clinical trials, including:

Downloaded from https://academic.oup.com/hmg/article-abstract/20/R1/R28/631150 by guest on 15 March 2019


using retrovirus vector-transduced fibroblast grafts and
delivery or upregulation of neurotrophic factors, generation
demonstrated restoration and survival of cholinergic neurons
of endogenous dopamine and alterations in neuronal circuitry,
in lesioned rodents and aged non-human primates. These
as well as implantation of supportive cells and dopaminergic
initial studies set the stage for the first phase I clinical trial
neurons. Gene delivery has typically been carried out using
of ex vivo NGF gene therapy (ClinicalTrials.gov Identifier:
AAV vectors with CED injection into the striatum. Vectors
NCT00017940). The rate of cognitive decline was slowed,
have delivered both glial-derived neurotrophic factor
and no apparent detrimental effects were observed arising
(GDNF; 87) and its analogue, neurturin (88), which serve as
from NGF expression 22 months post-engraftment (69).
trophic factors for dopaminergic neurons. Since abnormally
More recently, Ceregene has conducted phase I and II clinical
high GDNF can have toxic effects, efforts have gone into reg-
trials using an AAV vector that expresses NGF (Clinical-
ulating its expression through a tetracycline analogue system
Trials.gov Identifiers: NCT00087789 and NCT00876863,
(89) and by using highly specific zinc finger transcription
respectively). While the phase I trial demonstrated that stereo-
factors to upregulate expression of the endogenous gene,
tactic infusions of an NGF-expressing AAV vector is well tol-
with the latter imposing a physiologically controlled upper
erated, it is too early to know whether this gene therapy-based
limit (90).
strategy will significantly impact the course and symptomol-
In terms of modulating neurotransmitter pathways, all
ogy of the disease, given the phase II trial is currently ongoing.
three biosynthetic enzymes for dopamine have been included
During the past several years, promising efforts have
in the same lentivirus vector, including tyrosine hydroxylase
focused on reducing the levels of neurotoxic Ab peptide
(TH) and aromatic amino acid decarboxylase (AADC), as
species within the brain through removal of pathogenic Ab
well as GTP cyclohydrolase for synthesis of the biopterin
peptides or halting the proteolytic release of the amyloido-
co-factor for TH, with the assumption that dopamine, as a
genic form of Ab arising from pathogenic amyloid precursor
paracrine neurotransmitter, can be made by any cell in the
protein (APP) processing. Such preclinical vector-based thera-
striatum (91). However, serotonergic or GABAergic
pies include active and passive vaccines directed against
neurons can also produce dopamine as a false transmitter
specific epitopes of pathogenic Ab peptides (70– 75),
leading to dyskinesia (92). AAV-mediated delivery of
expression of Ab proteases (75– 78), delivery of small inhibi-
AADC alone has the advantage of making production of
tory RNAs (RNAi) designed to specifically suppress the
dopamine dependent on L-dopa intake, with ongoing phase
expression of APP processing enzymes (79 –82) and delivery
I trials indicating continued expression of AADC for at
of a cholesterol degrading enzyme into the brain (83). Each
least 2 years in the human brain (93). Investigators have
experimental approach has exhibited strong preclinical effi-
also tried to block hyperactive neurotransmission in PD
cacy in rodent models of AD, hence increasing enthusiasm
brains using an AAV vector to deliver the enzyme glutamic
for eventual translation of one or more of these strategies to
acid decarboxylase for synthesis of the inhibitory neurotrans-
clinical testing. The common challenge shared by gene thera-
mitter GABA into the subthalamic nucleus (94). Cell thera-
peutic approaches for AD which require intraparenchymal
pies have included a number of cell types, such as human
delivery relates to sufficiency of brain tissue coverage.
fetal mesencephalic tissue, which includes precursors of
Given AD impacts a number of human brain sub-regions
dopaminergic and serotonergic neurons, with the current
that are integral to learning and memory, viral vector strategies
focus on human dopaminergic neuroblasts generated from
must include the means to monitor vector infusion in real time
stem cells, in particular induced pluripotent stem (iPS)
to widely and safely disseminate vector particles within
cells derived from adult tissue (95).
afflicted networks. Recent advances in real-time monitoring
of CED-mediated AAV vector infusions within the non-human Strangely, few of these therapies address the etiology of the
primate brain indicate that this delivery hurdle can be disease based on genetic insights, and in fact most animal
overcome (84). models of PD employ lesioning of dopaminergic neuronal
connections, with the exception of transgenic mice overex-
pressing alpha-synuclein and recently developed models of
other PD genetic syndromes (96). Since three copies of the
Parkinson’s disease alpha-synuclein gene alone can cause PD (97) and upregulated
PD is a common progressive neurodegenerative disease. While alpha-synuclein inhibits neurotransmitter release (98), a
many regions of the brain are affected, the primary motor logical approach would be to decrease alpha-synuclein syn-
symptoms are caused by the loss of dopaminergic neurons in thesis with RNAi (84) or increase its degradation with the
Human Molecular Genetics, 2011, Vol. 20, Review Issue 1 R33

ubiquitin ligase, parkin (99). However, both the increased axonal transport (114). Also, misfolded SOD1 mutants
and decreased levels alpha-synuclein can cause neurodegen- have been shown to directly bind and inhibit mitochondrial
eration (84). voltage-dependent anion channel (VDAC1) leading to mito-
chondrial dysfunction (115). Taken together, these results
suggest that conformational changes in SOD1 (and possibly
Spinal muscular atrophy also other fALS-associated proteins), caused by mutations
SMA is an autosomal recessive disease caused by the loss of associated with some forms of fALS, and other as yet undeter-
function of the ‘survival of motor neuron’ gene, SMN1, mined factors, may be the basis for fALS and sALS. Gene
which leads to degeneration of motor neurons and infant mor- therapy approaches for ALS have centered on delivery of
tality. One approach to gene therapy would be to replace the growth factors such as IGF-1 (116– 118) and vascular endo-
missing gene in multiple motor neurons, which may now be thelial growth factor (116,119), and RNAi-based silencing of
possible with i.v. administration of AAV9 vectors which mutant SOD1 alleles (120 – 123) in transgenic ALS mouse

Downloaded from https://academic.oup.com/hmg/article-abstract/20/R1/R28/631150 by guest on 15 March 2019


were able to deliver the SMN1 cDNA to the spinal cord in a models. The most commonly used ALS mouse model carries
mouse model of SMA with marked correction of motor func- a human SOD1 G93A transgene expressed at high levels
tion and a dramatic increase in survival (100,101). However, with a mean survival of 129 days. To date, many of the
scaling delivery from mice to humans remains a huge chal- gene therapy studies have demonstrated a significant increase
lenge. Other strategies have taken advantage of the presence in median survival, but ultimately all treated mice have suc-
of a second copy of this gene, SMN2, coding for an identical cumbed to disease progression. The reasons for the apparent
amino acid sequence in the human genome (102). The failure of those interventions can be multifactorial, but an
SMN2 gene is defective due to a point mutation in an intron interesting aspect to consider is the fact that silencing
which interferes with correct splicing, such that a truncated, mutant SOD1 expression in motor neurons delays disease
non-functional protein is produced. In one therapeutic onset but not progression, while silencing it in microglia
modality, an ASO homologous to the last translated exon in does the opposite (124). This suggests that other CNS cell
SMN2 is used to guide an exonic splicing enhancer sequence types contribute significantly to the disease phenotype. Devel-
to the region as a binding platform for pre-mRNA splicing opment of new mouse models using mutant alleles from other
factors (103). This strategy gives increased SMN2 expression ALS genes will enhance our understanding of common disease
when infused into the lateral cerebral ventricles in an SMA pathways and development of effective gene therapies for this
mouse model (104,105). In another modality, transplicing is disease. The recent findings on the toxicity of misfolded SOD1
utilized in which the ASO binds to the defective intronic proteins (normal or mutant) strongly suggest that either immu-
sequence and carries a splicing domain and an intact final nization (125), or passive immunization with conformation-
exon from SMN1, with injection into the cerebral ventricles specific antibodies (126) may also be viable approaches to
also extending survival in a mouse model of SMA (106). develop effective therapies for ALS.

Amyotrophic lateral sclerosis Brain tumors


ALS is characterized by progressive muscle weakness result- Malignant glioma tumors (glioblastoma; GBM) are the most
ing from the loss of motor neurons in the brain and spinal common adult brain tumor and are virtually untreatable,
cord. Mutations in the SOD1 gene encoding cytosolic Cu/Zn with most patients dying within 2 years of diagnosis in spite
superoxide dismutase were the first to be linked to familial of neurosurgery, radiation and chemotherapy. These tumors
forms of ALS (fALS), 10% of all cases. In recent years, are very invasive in the brain and genetically heterogeneous
there has been a dramatic increase in the number of genes with a cancer stem cell population that defies all current che-
shown to be associated with fALS, such as ANG encoding motherapies (127). Some benign tumors of the nervous
angiogenin (107), TARDP encoding transactive response system, e.g. vestibular schwannomas, regress in response to
(TAR) DNA-binding protein TDP-43 (108), FUS encoding anti-angiogenic therapy (128), while for GBM tumors this
fused in sarcoma protein (also known as TLS for translocated treatment can enhance invasiveness without suppressing pro-
in liposarcoma) (109,110) and, more recently, optineurin liferation (129).
(111). Several other genes have been associated with rare Treatment of malignant brain tumors presents a major thera-
forms of fALS (reviewed in 112). The disease mechanism(s) peutic challenge requiring multicombinatorial approaches,
remains elusive. However, the fact that many fALS cases with gene/cell therapy showing promise as adjunct therapies.
are autosomal dominant suggests that disease-associated Experimental therapies in brain tumor models have focused
mutations generate protein species with toxic functions on direct elimination of tumor cells, including a ‘bystander’
instead of simple loss of function. An emerging picture in effect to confer selective toxicity to non-transduced tumor
the field is that some of the fALS-associated proteins may cells in the vicinity and on targeting invasive tumor cells.
also be involved in sporadic ALS (sALS) cases. As an The basic strategy is to remove the bulk of the tumor mass,
example, recent work has shown that oxidized wild-type and during that intervention to inject virus vectors or cells
SOD1 shares a conformational epitope with fALS-associated into the resected region to kill remaining tumor cells over an
mutant SOD1 (113), and can be found in spinal cord motor expanded radius. Concepts employed in this arsenal include:
neurons of a subset of sALS patients (114). Moreover, oxi- (i) Prodrug activation (or suicide genes). In this case, viral
dized wild-type SOD1 and mutant SOD1 share toxic proper- vectors or cells are used to express enzymes, such as viral thy-
ties to neurons by inhibition of kinesin-dependent fast midine kinase (130), bacterial cytosine deaminase/uracil
R34 Human Molecular Genetics, 2011, Vol. 20, Review Issue 1

phosphoribosyltransferase (131) or mammalian cytochrome using single-stranded DNA (73 bp; 152). In addition, there is
P450/carboxyesterase (132) within the brain to activate pro- the potential to deliver human artificial mini-chromosomes
drugs (ganciclovir, 5-fluorocytosine, cyclophosphamide/irino- into the cell nucleus for gene-deficiency syndromes
tecan, respectively) that can pass the BBB and be converted (153,154). These strategies include control of transplicing,
into active chemotherapeutic agents within the tumor. (ii) for example in SMA therapies (see above). Naturally occur-
Viral oncolysis. In this approach mutant forms of viruses, typi- ring and synthetic transposable elements, including the Sleep-
cally HSV-1 and adenovirus, are used which replicate selec- ing Beauty transposon, have shown promise in stable
tively in tumor versus normal brain based on cancer-related integration and gene expression in cells, including neurons
mutations or their increased proliferation rate (133). These in vivo, and are discussed in more detail elsewhere in this
oncolytic vectors can also be armed with therapeutic genes. issue. All these more ‘advanced’ manipulations of the
An increasing number of different viruses are being tested in genome will depend on large vector capacity and more effi-
this context, including measles virus, reovirus, Newcastle cient delivery to the nervous system. A new arena will be

Downloaded from https://academic.oup.com/hmg/article-abstract/20/R1/R28/631150 by guest on 15 March 2019


disease virus (134), polio virus (135) and vaccinia virus modulation of miRNA levels which have such an important
(136). Although promising, this field is still wrestling with role in development and cancer, including decreasing levels
immune responses to the virus and an unfavorable tumor of oncogenic miRNAs and increasing levels of tumor suppres-
microenvironment which restrict virus spread within the sor miRNAs (155). Epigenetic modulation of the genome is an
tumor (134). (iii) Cellular delivery. Several studies have expanding arena, with for example histone deacetylase inhibi-
shown that some normal cells introduced into the brain, such tors being used to slow down degeneration in mouse models of
as neural stem cells (137) and mesenchymal stem cells (138) Huntington’s disease (156).
migrate towards tumor foci. These cells can deliver agents
which are selectively toxic to tumor cells, e.g. antibodies
against tumor antigens (139,140), oncolytic virus vectors Improved vectors
(141,142), apoptotic factors (143) and anti-angiogenic proteins The new breed of viral vectors that can reach the CNS after
(144), as well as prodrug activating enzymes. (iv) Immu- peripheral administration (14,16) also transduce peripheral
notherapy to target tumor antigens. Several strategies have organs at high efficiency, namely the liver. This potential
been used to increase recognition of tumor antigens and limitation could be simply addressed with the use of
empower the immune system. These include vaccination CNS-specific promoters such as the synapsin-1, neuron-
with a common and unique GBM antigen, EGFRvIII (145) specific enolase or glial fibrillary acidic protein promoters.
and co-treatment with oncolytic HSV vectors and cyclopho- However, some of these promoters are relatively weak and
sphamide to temporarily suppress the immune curtailment of some are rather large, thus reducing considerably the trans-
virus spread, and inclusion of immune enhancing cytokines gene capacity of the vectors. An alternative approach is to
(146). In addition, T cells have been modified to express chi- incorporate into the transgene expression cassette targets for
meric receptors targeting antigens, such as the IL13 receptor miRNAs expressed at high levels in peripheral organs but
which is abundant on GBM tumors (147). (v) Zone of resist- absent or expressed at low levels in the CNS to prevent vector-
ance. Other efforts have tried to increase the resistance of mediated gene expression outside of the CNS (157, see
the brain to tumor invasion, including the use of AAV below). This approach appears to be quite universal to
vectors to infect normal brain cells so they release interferon- de-target vector-mediated expression from specific cell
beta which depresses tumor growth (148). lineages (57,158,159). Another approach to addressing the
undesired targeting of peripheral organs consists in engineer-
ing the viral vector particles themselves. Lentivirus vectors
FUTURE INITIATIVES can be pseudotyped with envelope proteins derived from
other viruses, and the vesicular stomatitis virus glycoprotein
New genetic manipulation methods is the most commonly used due to its pantropism. Previous
A number of new genetic strategies are being explored to studies have shown that incorporation of the rabies glyco-
manipulate the endogenous cell genome in order to regulate protein allows lentivirus vectors administered peripherally to
expression or correct mutated gene transcripts. In one reach the CNS via retrograde axonal transport (119,160). Non-
modality, zinc finger proteins (ZFPs) which recognize a typi- integrating lentivirus vectors can also be used effectively in
cally 18 bp site in the promoter region of a gene are linked to a the neural cells, thus reducing the risk of oncogenic insertion
transcriptional activator (149) to selectively turn on a gene, as events (161).
in the case of SMN2 (above; 90). This approach can also be Development of CNS-targeted AAV vectors has followed
used to stimulate gene correction by linking the ZFP to a three different routes: (i) in recent years hundreds of new
nuclease causing double-stranded breaks in the DNA in com- AAV capsids have been identified from humans and quite a
bination with a correct extrachromosomal DNA sequence such few other species, and systematic screening for new
that homologous recombination is stimulated (149). Another CNS-targeting properties may yet yield new AAV pseudo-
potent tool that can be linked to ZFPs are meganucleases types more powerful than AAV9; (ii) chimeric AAV2
which recognize .12 bp sequences and have been shown to capsids carrying CNS-targeting peptides identified from in
inactivate HIV sequences (150). Other methods to facilitate vivo phage display experiments can increase delivery. This
gene correction have included triplex-forming oligonucleo- approach had met with limited success until recently Chen
tides which can be used to modify sequence or inhibit gene et al. (17) showed that it can be used to develop AAV
expression (151). Gene correction has also been achieved vectors highly specific for brain microcapillary endothelial
Human Molecular Genetics, 2011, Vol. 20, Review Issue 1 R35

Downloaded from https://academic.oup.com/hmg/article-abstract/20/R1/R28/631150 by guest on 15 March 2019


Figure 3. Schematic representation of existing cellular therapies, including iPS cells, which require vector-based strategies to generate neuroprecursor cells and
neurons to ultimately treat neurodegenerative diseases. A variety of cellular therapies have been devised to repair and replace degenerated neuronal networks
within the CNS. These strategies utilize multiple sources of neurons and neuroprecursor cells, including fetal brain, pre-implantation embryos, mesenchymal
stem cells and iPS cells. The generation of iPS cells requires the use of viral vectors to deliver multiple genes key for the molecular reprogramming of
patient fibroblasts. These iPS cells can be differentiated into neurons or neuroprecursor cells that are subsequently transplanted into the diseased brain.

cells that are part of the BBB; (iii) molecular evolution of carrying genomes with mammalian expression cassettes
AAV vectors has been propelled by using new AAV capsid flanked by AAV2 inverted terminal repeats (169). These
libraries developed by DNA shuffling of existing AAV AAVP vectors appear to be quite effective at targeting brain
capsid genes (162– 164). This approach has been used to gen- tumor vasculature (170).
erate new AAV vectors with improved transduction properties The recent developments in engineering new CNS-targeted
for particular targets, such as lung epithelium (165), myocar- vectors give great optimism that within the next decade we
dium (166) and more recently to a particular subset of will finally achieve the long-standing goal of global gene
neurons in the piriform cortex in the brain following delivery to the post-natal CNS via the vasculature. This will
seizure-induced temporary disruption of the BBB (167). This undoubtedly revolutionize neuroscience research, and mark a
selection is performed in live animals, and one of the highly new era in neurology with the genetic tools existing to
significant findings has been newly identified chimeric AAV develop effective therapies for many conditions that remain
capsids with dramatically reduced transduction of liver upon untreatable today. But before then, we already have excep-
systemic infusion. The work by Kumar et al. (168) where a tional tools at hand to change the outcome of many neurologi-
rabies virus glycoprotein-derived peptide was used to target cal diseases using viral vectors for focal gene delivery to
an artificial siRNA to the CNS suggests that the distinction particular structures in the CNS. Most work thus far has
between artificial/synthetic and viral vectors is likely to relied on the use of strong constitutive promoters, but in
become increasingly blurred as both sides borrow from each many instances it may be necessary to regulate gene
other to achieve the ultimate goal. AAV-phage (AAVP) expression. This has been accomplished to a large extent by
vectors are another example of this melding of components incorporating into the gene transfer vectors a combination
with target peptide-displaying M13-derived phage capsids of drug-responsive transcription factors and respective
R36 Human Molecular Genetics, 2011, Vol. 20, Review Issue 1

promoters, thus being able to regulate transgene expression in Conflict of Interest statement. None declared.
the CNS by peripheral administration of a particular drug.
The most commonly used and investigated systems for
this purpose are the tetracycline-responsive (89,171– 173)
and rapamycin-responsive systems (174,175). These FUNDING
drug-regulated viral vector systems have yet to be tested in Support for X.O.B. comes from NIH/NCI CA069246 and NIH/
the CNS of patients. NINDS NS242791, for W.J.B. from NIH R01-AG026328 and
for M.S.-E. from NIH/NINDS R01NS066310, U01NS064096
and NIH/NICHHD R01HD060576.
IPs and stem cells
Neuronal cell death is associated with most of the major dis-

Downloaded from https://academic.oup.com/hmg/article-abstract/20/R1/R28/631150 by guest on 15 March 2019


eases that afflict the CNS. Stem cell therapeutics holds REFERENCES
promise for replacing degenerating or ablated neurons to ulti-
1. Abbott, N.J., Ronnback, L. and Hansson, E. (2006) Astrocyte-endothelial
mately restore neuronal network integrity. While embryonic interactions at the blood-brain barrier. Nat. Rev. Neurosci., 7, 41– 53.
stem cell-based therapy has been approved for clinical trial 2. Armulik, A., Genové, G., Mäe, M., Nisancioglu, M.H., Wallgard, E.,
testing in patients with spinal cord lesions (176), the ethical Niaudet, C., He, L., Norlin, J., Lindblom, P., Strittmatter, K. et al. (2010)
implications and funding policy inconsistency associated Pericytes regulate the blood-brain barrier. Nature, 468, 557– 561.
with the use of stem cells isolated from human embryos for 3. Bobo, R.H., Laske, D.W., Akbasak, A., Morrison, P.F., Dedrick, R.L.
and Oldfield, E.H. (1994) Convection-enhanced delivery of
CNS therapy have led to the rapid development of new cell macromolecules in the brain. Proc. Natl Acad. Sci. USA, 91, 2076–2080.
sources, including iPS cells (Fig. 3). iPS cells require the con- 4. Salegio, E.A., Kells, A.P., Richardson, R.M., Hadaczek, P., Forsayeth, J.,
trolled ‘reprogramming’ of adult somatic cells through careful Bringas, J., Sardi, S.P., Passini, M.A., Shihabuddin, L.S., Cheng, S.H.
introduction of key molecular cues and re-derivation of viable et al. (2010) Magnetic resonance imaging-guided delivery of
adeno-associated virus type 2 to the primate brain for the treatment of
clones that lack tumorigenic potential (177,178). Viral vectors, lysosomal storage disorders. Hum. Gene Ther., 21, 1093–1103.
including those derived from retrovirus and lentivirus, have 5. Passini, M.A., Watson, D.J., Vite, C.H., Landsburg, D.J., Peigenbaum,
proved useful for efficient delivery of genes encoding these A.L. and Wolfe, J.H. (2003) Intraventricular brain injection of
molecular cues (Oct3/4, Sox, Klf, Myc, Nanog and LIN28) adeno-associated virus type 1 (AAV1) in neonatal mice results in
to adult human fibroblasts (179 – 181). The challenges facing complementary patterns of neuronal transduction to AAV2 and total
long-term correction of storage lesions in the brains of
the clinical implementation of iPS cells for neurodegenerative beta-glucuronidase-deficient mice. J. Virol., 77, 7034– 7040.
diseases relate to achieving an optimal balancing of repro- 6. Watson, G., Bastacky, J., Belichenko, P., Buddhikot, M., Jungles
gramming factors [reviewed by Lewitzky and Yamanaka S., Vellard, M., Mobley, W.C. and Kakkis, E. (2006) Intrathecal
(182)], lowering risk of integrating vector-mediated gene dis- administration of AAV vectors for the treatment of lysosomal storage in
the brains of MPS I mice. Gene Ther., 13, 917– 925.
ruption (183,184) and prevention of teratoma formation, 7. Cearley, C.N., Vandenberghe, L.H., Parente, M.K., Carnish, E.R.,
which is an inherent risk of iPS cell derivation (185,186). Wilson, J.M. and Wolfe, J.H. (2008) Expanded repertoire of AAV vector
serotypes mediate unique patterns of transduction in mouse brain. Mol.
Ther., 16, 1710–1718.
8. Liu, G., Martins, I.H., Chiorini, J.A. and Davidson, B.L. (2005)
State-of-the-art summary Adeno-associated virus type 4 (AAV4) targets ependyma and astrocytes
In the past 20 years, genetic therapy has moved from a dream in the subventricular zone and RMS. Gene Ther., 12, 1503– 1508.
9. Meijer, D.H., Maguire, C.A., LeRoy, S.G. and Sena-Esteves, M. (2009)
to a near medical reality for some diseases, with marked ben- Controlling brain tumor growth by intraventricular administration of an
eficial effects on immune function in X-linked SCID cases AAV vector encoding IFN-beta. Cancer Gene Ther., 16, 664– 671.
(187) and restoration of some vision in Leber’s optic 10. Dickson, P., McEntee, M., Vogler, C., Le, S., Levy, B., Peinovich, M.,
atrophy (188). There have been many hurdles along the way, Hanson, S., Passage, M. and Kakkis, E. (2007) Intrathecal enzyme
replacement therapy: successful treatment of brain disease via the
including toxicity of vectors, oncogenic insertions into the cerebrospinal fluid. Mol. Genet. Metab., 91, 61–68.
genome, immune rejections and difficulties in scaling up 11. Chang, M., Cooper, J.D., Sleat, D.E., Cheng, S.H., Dodge, J.C., Passini,
from mouse models to humans. The burgeoning successes M.A., Lobel, P. and Davidson, B.L. (2008) Intraventricular enzyme
have depended in large part on genetics, including identifying replacement improves disease phenotypes in a mouse model of late
disease genes and types of mutations, design of infantile neuronal ceroid lipofuscinosis. Mol. Ther., 16, 649– 656.
12. Pardridge, W.M. (2010) Biopharmaceutical drug targeting to the brain.
genome-interacting elements, manipulation of splicing events J. Drug Target, 18, 157 –167.
and novel genetic elements, such as microRNAs and regulat- 13. Osborn, M.J., McElmurry, R.T., Peacock, B., Tolar, J. and Blazar
ory non-coding RNAs. Critical factors lie in tailoring the B.R. (2008) Targeting of the CNS in MPS-IH using a nonviral
mode and content of the genetic vehicle to the specifics of transferrin-alpha-L-iduronidase fusion gene product. Mol. Ther.,
16, 1459– 1466.
the disease, and effectively utilizing the broad range of target- 14. Foust, K.D., Nurre, E., Montgomery, C.L., Hernandez, A., Chan, C.M.
ing modalities and vector types available. and Kaspar, B.K. (2009) Intravascular AAV9 preferentially targets
neonatal neurons and adult astrocytes. Nat. Biotechnol., 27, 59–65.
15. Duque, S., Joussemet, B., Riviere, C., Marais, T., Dubreil, L., Douar,
A.M., Fyfe, J., Moullier, P., Colle, M.A. and Barkats, M. (2009)
ACKNOWLEDGEMENTS Intravenous administration of self-complementary AAV9 enables
transgene delivery to adult motor neurons. Mol. Ther., 17, 1187– 1196.
We thank Ms Suzanne McDavitt for skilled editorial assist- 16. Louboutin, J.P., Chekmasova, A.A., Marusich, E., Chowdhury, J.R. and
ance, and Ms Emily Mills at Millstone Design (http Strayer, D.S. (2010) Efficient CNS gene delivery by intravenous
://www.millstone-design.com) for preparation of figures. injection. Nat. Methods, 7, 905– 907.
Human Molecular Genetics, 2011, Vol. 20, Review Issue 1 R37

17. Chen, Y.H., Chang, M. and Davidson, B.L. (2009) Molecular signatures vectors in the brains of animals with pre-existing anti-adenoviral
of disease brain endothelia provide new sites for CNS-directed enzyme immunity: clinical implications. Mol. Ther., 15, 2154–2163.
therapy. Nat. Med., 15, 1215–1218. 35. Oehmig, A., Fraefel, C. and Breakefield, X.O. (2004) Update of
18. Biffi, A., De Palma, M., Quattrini, A., Del Carro, U., Amadio, S., herpesvirus amplicon vectors. Mol. Ther., 10, 630– 643.
Visigalli, I., Sessa, M., Fasano, S., Brambilla, R., Marchesini, S. et al. 36. Peruzzi, P.P., Lawler, S.E., Senior, S.L., Dmitrieva, N., Edser, P.A.,
(2004) Correction of metachromatic leukodystrophy in the mouse model Gianni, D., Chiocca, E.A. and Wade-Martins, R. (2009) Physiological
by transplantation of genetically modified hematopoietic stem cells. transgene regulation and functional complementation of a neurological
J. Clin. Invest., 113, 1118–1129. disease gene deficiency in neurons. Mol. Ther., 17, 1517–1526.
19. De Palma, M., Mazzieri, R., Politi, L.S., Pucci, F., Zonari, E., Sitia, G., 37. Lufino, M.M., Edser, P.A. and Wade-Martins, R. (2008) Advances in
Mazzoleni, S., Moi, D., Venneri, M.A., Indraccolo, S. et al. (2008) high-capacity extrachromosomal vector technology: episomal
Tumor-targeted interferon-alpha delivery by Tie2-expressing monocytes maintenance, vector delivery, and transgene expression. Mol. Ther., 16,
inhibits tumor growth and metastasis. Cancer Cell, 14, 299–311. 1525– 1538.
20. Aboody, K.S., Brown, A., Rainov, N., Bower, K., Liu, S., Yang, W., 38. Hickman, S., Shapiro, L.J. and Neufeld, E.F. (1974) A recognition
Small, J., Herrlinger, U., Ourednik, V., Black, P. et al. (2000) Neural marker required for uptake of a lysosomal enzyme by cultured

Downloaded from https://academic.oup.com/hmg/article-abstract/20/R1/R28/631150 by guest on 15 March 2019


stem cells display extensive tropism for pathology in adult brain: fibroblasts. Biochem. Biophys. Res. Commun., 57, 55– 61.
evidence from intracranial gliomas. Proc. Natl Acad. Sci. USA, 97, 39. Skorupa, A.F., Fisher, K.J., Wilson, J.M., Parente, M.K. and Wolfe, J.H.
12846–12851. (1999) Sustained production of beta-glucuronidase from localized sites
21. Stitelman, D.H., Endo, M., Bora, A., Muvarak, N., Zoltick, P.W., Flake, after AAV vector gene transfer results in widespread distribution of
A.W. and Brazelton, T.R. (2010) Robust in vivo transduction of nervous enzyme and reversal of lysosomal storage lesions in a large volume of
system and neural stem cells by early gestational intra amniotic gene brain in mucopolysaccharidosis VII mice. Exp. Neurol., 160, 17–27.
transfer using lentiviral vector. Mol. Ther., 18, 1615–1623. 40. Bosch, A., Perret, E., Desmaris, N. and Heard, J.M. (2000) Long-term and
22. Zhang, Y., Calon, F., Zhu, C., Boado, R.J. and Pardridge, W.M. (2003b) significant correction of brain lesions in adult mucopolysaccharidosis type
Intravenous nonviral gene therapy causes normalization of striatal VII mice using recombinant AAV vectors. Mol. Ther., 1, 63– 70.
tyrosine hydroxylase and reversal of motor impairment in experimental 41. Bosch, A., Perret, E., Desmaris, N., Trono, D. and Heard, J.M. (2000)
parkinsonism. Hum. Gene Ther., 14, 1– 12. Reversal of pathology in the entire brain of mucopolysaccharidosis type
23. Stachowiak, E.K., Roy, I., Lee, Y.W., Capacchietti, M., Aletta, J.M., VII mice after lentivirus-mediated gene transfer. Hum. Gene Ther., 11,
Prasad, P.N. and Stachowiak, M.K. (2009) Targeting novel integrative 1139– 1150.
nuclear FGFR1 signaling by nanoparticle-mediated gene transfer 42. Consiglio, A., Quattrini, A., Martino, S., Bensadoun, J.C., Dolcetta, D.,
stimulates neurogenesis in the adult brain. Integr. Biol. (Camb.), 1, Trojani, A., Benaglia, G., Marchesini, S., Cestari, V., Oliverio, A. et al.
394– 403. (2001) In vivo gene therapy of metachromatic leukodystrophy by
24. De Vry, J., Martinez-Martinez, P., Losen, M., Temel, Y., Steckler, T., lentiviral vectors: correction of neuropathology and protection against
Steinbusch, H.W., De Baets, M.H. and Prickaerts, J. (2010) In vivo
learning impairments in affected mice. Nat. Med., 7, 310–316.
electroporation of the central nervous system: a non-viral approach for
43. Passini, M.A., Macauley, S.L., Huff, M.R., Taksir, T.V., Bu, J., Wu, I.H.,
targeted gene delivery. Prog. Neurobiol., 92, 227–244.
Piepenhagen, P.A., Dodge, J.C., Shihabuddin, L.S., O’Riordan, C.R.
25. Nagayama, S., Zeng, S., Xiong, W., Fletcher, M.L., Masurkar, A.V.,
et al. (2005) AAV vector-mediated correction of brain pathology in a
Davis, D.J., Pieribone, V.A. and Chen, W.R. (2007) In vivo
mouse model of Niemann-Pick A disease. Mol. Ther., 11, 754– 762.
simultaneous tracing and Ca(2+) imaging of local neuronal circuits.
44. Passini, M.A., Bu, J., Fidler, J.A., Ziegler, R.J., Foley, J.W., Dodge, J.C.,
Neuron, 53, 789– 803.
Yang, W.W., Clarke, J., Taksir, T.V., Griffiths, D.A. et al. (2007)
26. Nevian, T. and Helmchen, F. (2007) Calcium indicator loading of
Combination brain and systemic injections of AAV provide maximal
neurons using single-cell electroporation. Pflugers Arch., 454, 675–688.
27. Chesler, A.T., Le Pichon, C.E., Brann, J.H., Araneda, R.C., Zou, D.J. and functional and survival benefits in the Niemann-Pick mouse. Proc. Natl
Firestein, S. (2008) Selective gene expression by postnatal Acad. Sci. USA, 104, 9505–9510.
electroporation during olfactory interneuron neurogenesis. PLoS ONE, 45. Passini, M.A., Dodge, J.C., Bu, J., Yang, W., Zhao, Q., Sondhi, D.,
3, e1517. Hackett, N.R., Kaminsky, S.M., Mao, Q., Shihabuddin, L.S. et al. (2006)
28. Hadaczek, P., Eberling, J.L., Pivirotto, P., Bringas, J., Forsayeth, J. and Intracranial delivery of CLN2 reduces brain pathology in a mouse model
Bankiewicz, K.S. (2010) Eight years of clinical improvement in of classical late infantile neuronal ceroid lipofuscinosis. J. Neurosci.,
MPTP-lesioned primates after gene therapy with AAV2-hAADC. Mol. 26, 1334–1342.
Ther., 18, 1458–1461. 46. Davidson, B.L., Stein, C.S., Heth, J.A., Martins, I., Kotin, R.M.,
29. Kaplitt, M.G., Feigin, A., Tang, C., Fitzsimons, H.L., Mattis, P., Lawlor, Derksen, T.A., Zabner, J., Ghodsi, A. and Chiorini, J.A. (2000)
P.A., Bland, R.J., Young, D., Strybing, K., Eidelberg, D. et al. (2007) Recombinant adeno-associated virus type 2, 4, and 5 vectors:
Safety and tolerability of gene therapy with an adeno-associated virus transduction of variant cell types and regions in the mammalian central
(AAV) borne GAD gene for Parkinson’s disease: an open label, phase I nervous system. Proc. Natl Acad. Sci. USA, 97, 3428–3432.
trial. Lancet, 369, 2097–2105. 47. Taylor, R.M. and Wolfe, J.H. (1997) Decreased lysosomal storage in the
30. Worgall, S., Sondhi, D., Hackett, N.R., Kosofsky, B., Kekatpure, M.V., adult MPS VII mouse brain in the vicinity of grafts of retroviral
Neyzi, N., Dyke, J.P., Ballon, D., Heier, L., Greenwald, B.M. et al. vector-corrected fibroblasts secreting high levels of beta-glucuronidase.
(2008) Treatment of late infantile neuronal ceroid lipofuscinosis by CNS Nat. Med., 3, 771– 774.
administration of a serotype 2 adeno-associated virus expressing CLN2 48. Passini, M.A., Lee, E.B., Heuer, G.G. and Wolfe, J.H. (2002)
cDNA. Hum. Gene Ther., 19, 463– 474. Distribution of a lysosomal enzyme in the adult brain by axonal transport
31. Eberling, J.L., Jagust, W.J., Christine, C.W., Starr, P., Larson, P., and by cells of the rostral migratory stream. J. Neurosci., 22, 6437–
Bankiewicz, K.S. and Aminoff, M.J. (2008) Results from a phase I safety 6446.
trial of hAADC gene therapy for Parkinson disease. Neurology, 70, 49. Hennig, A.K., Levy, B., Ogilvie, J.M., Vogler, C.A., Galvin, N.,
1980– 1983. Bassnett, S. and Sands, M.S. (2003) Intravitreal gene therapy reduces
32. Christine, C.W., Starr, P.A., Larson, P.S., Eberling, J.L., Jagust, W.J., lysosomal storage in specific areas of the CNS in mucopolysaccharidosis
Hawkins, R.A., VanBrocklin, H.F., Wright, J.F., Bankiewicz, K.S. and VII mice. J. Neurosci., 23, 3302–3307.
Aminoff, M.J. (2009) Safety and tolerability of putaminal AADC gene 50. Dodge, J.C., Clarke, J., Song, A., Bu, J., Yang, W., Taksir, T.V.,
therapy for Parkinson disease. Neurology, 73, 1662– 1669. Griffiths, D., Zhao, M.A., Schuchman, E.H., Cheng, S.H. et al. (2005)
33. McPhee, S.W., Janson, C.G., Li, C., Samulski, R.J., Camp, A.S., Francis, Gene transfer of human acid sphingomyelinase corrects neuropathology
J., Shera, D., Lioutermann, L., Feely, M., Freese, A. et al. (2006) and motor deficits in a mouse model of Niemann-Pick type A disease.
Immune responses to AAV in a phase I study for Canavan disease. Proc. Natl Acad. Sci. USA, 102, 17822–17827.
J. Gene Med., 8, 577– 588. 51. Cearley, C.N. and Wolfe, J.H. (2007) A single injection of an
34. Barcia, C., Jimenez-Dalmaroni, M., Kroeger, K.M., Puntel, M., adeno-associated virus vector into nuclei with divergent connections
Rapaport, A.J., Larocque, D., King, G.D., Johnson, S.A., Liu, C., Xiong, results in widespread vector distribution in the brain and global
W. et al. (2007) One-year expression from high-capacity adenoviral correction of a neurogenetic disease. J. Neurosci., 27, 9928–9940.
R38 Human Molecular Genetics, 2011, Vol. 20, Review Issue 1

52. Baek, R.C., Broekman, M.L., Leroy, S.G., Tierney, L.A., Sandberg, 71. Bowers, W.J., Mastrangelo, M.A., Stanley, H.A., Casey, A.E., Milo,
M.A., d’Azzo, A., Seyfried, T.N. and Sena-Esteves, M. (2010) L.J.J. and Federoff, H.J. (2005) HSV amplicon-mediated Abeta
AAV-mediated gene delivery in adult GM1-gangliosidosis mice corrects vaccination in Tg2576 mice: differential antigen-specific immune
lysosomal storage in CNS and improves survival. PLoS ONE, 5, e13468. responses. Neurobiol. Aging, 26, 393 –407.
53. Lattanzi, A., Neri, M., Maderna, C., di Girolamo, I., Martino, S., 72. Kim, H.D., Maxwell, J.A., Kong, F.K., Tang, D.C. and Fukuchi, K.
Orlacchio, A., Amendola, M., Naldini, L. and Gritti, A. (2010) (2005) Induction of anti-inflammatory immune response by an
Widespread enzymatic correction of CNS tissues by a single adenovirus vector encoding 11 tandem repeats of Abeta1– 6: toward
intracerebral injection of therapeutic lentiviral vector in leukodystrophy safer and effective vaccines against Alzheimer’s disease. Biochem.
mouse models. Hum. Mol. Genet., 19, 2208– 2227. Biophys. Res. Commun., 336, 84–92.
54. Vite, C.H., McGowan, J.C., Niogi, S.N., Passini, M.A., Drobatz, K.J., 73. Fukuchi, K., Tahara, K., Kim, H.D., Maxwell, J.A., Lewis, T.L.,
Haskins, M.E. and Wolfe, J.H. (2005) Effective gene therapy for an Accavitti-Loper, M.A., Kim, H., Ponnazhagan, S. and Lalonde, R. (2006)
inherited CNS disease in a large animal model. Ann. Neurol., 57, Anti-Abeta single-chain antibody delivery via adeno-associated virus for
355– 364. treatment of Alzheimer’s disease. Neurobiol. Dis., 23, 502–511.
55. Ciron, C., Desmaris, N., Colle, M.A., Raoul, S., Joussemet, B., Verot, L., 74. Levites, Y., Jansen, K., Smithson, L.A., Dakin, R., Holloway, V.M., Das,

Downloaded from https://academic.oup.com/hmg/article-abstract/20/R1/R28/631150 by guest on 15 March 2019


Ausseil, J., Froissart, R., Roux, F., Cherel, Y. et al. (2006) Gene therapy P. and Golde, T.E. (2006) Intracranial adeno-associated virus-mediated
of the brain in the dog model of Hurler’s syndrome. Ann. Neurol., 60, delivery of anti-pan amyloid beta, amyloid beta40, and amyloid beta42
204– 213. single-chain variable fragments attenuates plaque pathology in amyloid
56. Biffi, A., Capotondo, A., Fasano, S., del Carro, U., Marchesini, S., precursor protein mice. J. Neurosci., 26, 11923– 11928.
Azuma, H., Malaguti, M.C., Amadio, S., Brambilla, R., Grompe, M. 75. Ryan, D.A., Mastrangelo, M.A., Narrow, W.C., Sullivan, M.A.,
et al. (2006) Gene therapy of metachromatic leukodystrophy reverses Federoff, H.J. and Bowers, W.J. (2010) Abeta-directed single-chain
neurological damage and deficits in mice. J. Clin. Invest., 116, antibody delivery via a serotype-1 AAV vector improves learning
3070–3082. behavior and pathology in Alzheimer’s disease mice. Mol. Ther., 18,
57. Gentner, B., Visigalli, I., Hiramatsu, H., Lechman, E., Ungari, S., 1471– 1481.
Giustacchini, A., Schira, G., Amendola, M., Quattrini, A., Martino, S. 76. Marr, R.A., Rockenstein, E., Mukherjee, A., Kindy, M.S., Hersh, L.B.,
et al. (2010) Identification of hematopoietic stem cell-specific miRNAs Gage, F.H., Verma, I.M. and Masliah, E. (2003) Neprilysin gene transfer
enables gene therapy of globoid cell leukodystrophy. Sci. Transl. Med., reduces human amyloid pathology in transgenic mice. J. Neurosci., 23,
2, 58–84. 1992– 1996.
58. Visigalli, I., Delai, S., Politi, L.S., Di Domenico, C., Cerri, F., Mrak, E., 77. Iwata, N., Tsubuki, S., Takaki, Y., Shirotani, K., Lu, B., Gerard, N.P.,
D’Isa, R., Ungaro, D., Stok, M., Sanvito, F. et al. (2010) Gene therapy Gerard, C., Hama, E., Lee, H.J. and Saido, T.C. (2001) Metabolic
augments the efficacy of hematopoietic cell transplantation and fully regulation of brain Abeta by neprilysin. Science, 292, 1550–1552.
corrects mucopolysaccharidosis type I phenotype in the mouse model. 78. Lebson, L., Nash, K., Kamath, S., Herber, D., Carty, N., Lee, D.C., Li,
Blood, 116, 5130– 5139. Q., Szekeres, K., Jinwal, U., Koren, J. et al. (2010) Trafficking
59. Bartus, R.T., Dean, R.L., Beer, B. 3rd and Lippa, A.S. (1982) The CD11b-positive blood cells deliver therapeutic genes to the brain of
cholinergic hypothesis of geriatric memory dysfunction. Science, 217, amyloid-depositing transgenic mice. J. Neurosci., 30, 9651–9658.
408– 414. 79. Thinakaran, G. (1999) The role of presenilins in Alzheimer’s disease.
60. Coyle, J.T., Price, D.L. and DeLong, M.R. (1983) Alzheimer’s disease: a J. Clin. Invest., 104, 1321–1327.
disorder of cortical cholinergic innervation. Science, 219, 1184–1190. 80. Wolfe, M.S., Xia, W., Ostaszewski, B.L., Diehl, T.S., Kimberly, W.T.
61. Masliah, E., Terry, R.D., Alford, M., DeTeresa, R. and Hansen, L.A. and Selkoe, D.J. (1999) Two transmembrane aspartates in presenilin-1
(1991) Cortical and subcortical patterns of synaptophysinlike required for presenilin endoproteolysis and gamma-secretase activity.
immunoreactivity in Alzheimer’s disease. Am. J. Pathol., 138, 235– 246. Nature, 398, 513–517.
62. Perry, E.K., Perry, R.H., Blessed, G. and Tomlinson, B.E. (1978) 81. Singer, O., Marr, R.A., Rockenstein, E., Crews, L., Coufal, N.G., Gage,
Changes in brain cholinesterases in senile dementia of Alzheimer type. F.H., Verma, I.M. and Masliah, E. (2005) Targeting BACE1 with
Neuropathol. Appl. Neurobiol., 4, 273–277. siRNAs ameliorates Alzheimer disease neuropathology in a transgenic
63. Perry, E.K., Tomlinson, B.E., Blessed, G., Bergmann, K., Gibson, P.H. model. Nat. Neurosci., 8, 1343–1349.
and Perry, R.H. (1978) Correlation of cholinergic abnormalities with 82. Hong, C.S., Goins, W.F., Goss, J.R., Burton, E.A. and Glorioso, J.C.
senile plaques and mental test scores in senile dementia. Br. Med. J., 2, (2006) Herpes simplex virus RNAi and neprilysin gene transfer vectors
1457–1459. reduce accumulation of Alzheimer’s disease-related amyloid-beta
64. Blesch, A. and Tuszynski, M. (1995) Ex vivo gene therapy for peptide in vivo. Gene Ther., 13, 1068– 1079.
Alzheimer’s disease and spinal cord injury. Clin. Neurosci., 3, 268– 274. 83. Hudry, E., Van Dam, D., Kulik, W., De Deyn, P.P., Stet, F.S.,
65. Tuszynski, M.H. and Gage, F.H. (1995) Bridging grafts and transient Ahouansou, O., Benraiss, A., Delacourte, A., Bougnères, P., Aubourg, P.
nerve growth factor infusions promote long-term central nervous system et al. (2010) Adeno-associated virus gene therapy with cholesterol
neuronal rescue and partial functional recovery. Proc. Natl Acad. Sci. 24-hydroxylase reduces the amyloid pathology before or after the onset
USA, 92, 4621–4625. of amyloid plaques in mouse models of Alzheimer’s disease. Mol. Ther.,
66. Tuszynski, M.H. and Gage, F.H. (1990) Potential use of neurotrophic 18, 44–53.
agents in the treatment of neurodegenerative disorders. Acta Neurobiol. 84. Gorbatyuk, O.S., Li, S., Nash, K., Gorbatyuk, M., Lewin, A.S., Sullivan,
Exp. (Warsz.), 50, 311–322. L.F., Mandel, R.J., Chen, W., Meyers, C., Manfredsson, F.P. et al. (2010)
67. Tuszynski, M.H., Senut, M.C., Ray, J. and Roberts, J. (1994) Somatic In vivo RNAi-mediated alpha-synuclein silencing induces nigrostriatal
gene transfer to the adult primate central nervous system: in vitro and in degeneration. Mol. Ther., 18, 1450– 1457.
vivo characterization of cells genetically modified to secrete nerve 85. Shulman, J.M., De Jager, P.L. and Feany, M.B. (2011) Parkinson’s
growth factor. Neurobiol. Dis., 1, 67–78. disease: genetics and pathogenesis. Annu. Rev. Pathol., 6, 193–222.
68. Rosenberg, M.B., Friedmann, T., Robertson, R.C., Tuszynski, M., Wolff, 86. Klein, C. and Schlossmacher, M.G. (2007) Parkinson disease, 10 years
J.A., Breakefield, X.O. and Gage, F.H. (1988) Grafting genetically after its genetic revolution: multiple clues to a complex disorder.
modified cells to the damaged brain: restorative effects of NGF Neurology, 69, 2093– 2104.
expression. Science, 242, 1575–1578. 87. Kells, A.P., Eberling, J., Su, X., Pivirotto, P., Bringas, J., Hadaczek, P.,
69. Tuszynski, M.H., Thal, L., Pay, M., Salmon, D.P., U, H.S., Bakay, R., Narrow, W.C., Bowers, W.J., Federoff, H.J., Forsayeth, J. et al. (2010)
Patel, P., Blesch, A., Vahlsing, H.L., Ho, G. et al. (2005) A phase 1 Regeneration of the MPTP-lesioned dopaminergic system after
clinical trial of nerve growth factor gene therapy for Alzheimer disease. convection-enhanced delivery of AAV2-GDNF. J. Neurosci., 30,
Nat. Med., 11, 551– 555. 9567– 9577.
70. Zhang, J., Wu, X., Qin, C., Qi, J., Ma, S., Zhang, H., Kong, Q., Chen, D., 88. Marks, W.J.J., Bartus, R.T., Siffert, J., Davis, C.S., Lozano, A., Boulis,
Ba, S. and He, W. (2003a) A novel recombinant adeno-associated virus N., Vitek, J., Stacy, M., Turner, D., Verhagen, L. et al. (2010) Gene
vaccine reduces behavioral impairment and beta-amyloid plaques in a delivery of AAV2-neurturin for Parkinson’s disease: a double-blind,
mouse model of Alzheimer’s disease. Neurobiol. Dis., 14, 365– 379. randomised, controlled trial. Lancet Neurol., 9, 1164– 1172.
Human Molecular Genetics, 2011, Vol. 20, Review Issue 1 R39

89. Manfredsson, F.P., Burger, C., Rising, A.C., Zuobi-Hasona, K., Sullivan, (2006) ANG mutations segregate with familial and ‘sporadic’
L.F., Lewin, A.S., Huang, J., Piercefield, E., Muzyczka, N. and Mandel, amyotrophic lateral sclerosis. Nat. Genet., 38, 411– 413.
R.J. (2009) Tight long-term dynamic doxycycline responsive 108. Sreedharan, J., Blair, I.P., Tripathi, V.B., Hu, X., Vance, C., Rogelj, B.,
nigrostriatal GDNF using a single rAAV vector. Mol. Ther., 17, Ackerley, S., Durnall, J.C., Williams, K.L., Buratti, E. et al. (2008)
1857– 1867. TDP-43 mutations in familial and sporadic amyotrophic lateral sclerosis.
90. Laganiere, J., Kells, A.P., Lai, J.T., Guschin, D., Paschon, D.E., Meng, X., Science, 319, 1668–1672.
Fong, L.K., Yu, Q., Rebar, E.J., Gregory, P.D. et al. (2010) An engineered 109. Kwiatkowski, T.J., Bosco, D.A., Leclerc, A.L., Tamrazian, E.,
zinc finger protein activator of the endogenous glial cell line-derived Vanderburg, C.R., Russ, C., Davis, A., Gilchrist, J., Kasarskis, E.J.,
neurotrophic factor gene provides functional neuroprotection in a rat Munsat, T. et al. (2009) Mutations in the FUS/TLS gene on
model of Parkinson’s disease. J. Neurosci., 30, 16469–16474. chromosome 16 cause familial amyotrophic lateral sclerosis. Science,
91. Azzouz, M., Martin-Rendon, E., Barber, R.D., Mitrophanous, K.A., 323, 1205– 1208.
Carter, E.E., Rohll, J.B., Kingsman, S.M., Kingsman, A.J. and 110. Vance, C., Rogelj, B., Hortobágyi, T., De Vos, K.J., Nishimura, A.L.,
Mazarakis, N.D. (2002) Multicistronic lentiviral vector-mediated striatal Sreedharan, J., Hu, X., Smith, B., Ruddy, D., Wright, P. et al. (2009)
gene transfer of aromatic L-amino acid decarboxylase, tyrosine Mutations in FUS, an RNA processing protein, cause familial

Downloaded from https://academic.oup.com/hmg/article-abstract/20/R1/R28/631150 by guest on 15 March 2019


hydroxylase, and GTP cyclohydrolase I induces sustained transgene amyotrophic lateral sclerosis type 6. Science, 323, 1208–1211.
expression, dopamine production, and functional improvement in a rat 111. Maruyama, H., Morino, H., Ito, H., Izumi, Y., Kato, H., Watanabe, Y.,
model of Parkinson’s disease. J. Neurosci., 22, 10302– 10312. Kinoshita, Y., Kamada, M., Nodera, H., Suzuki, H. et al. (2010)
92. Muñoz, A., Carlsson, T., Tronci, E., Kirik, D., Björklund, A. and Carta, Mutations of optineurin in amyotrophic lateral sclerosis. Nature, 465,
M. (2009) Serotonin neuron-dependent and -independent reduction of 223–226.
dyskinesia by 5-HT1A and 5-HT1B receptor agonists in the rat 112. Dion, P.A., Daoud, H. and Rouleau, G.A. (2009) Genetics of motor
Parkinson model. Exp. Neurol., 219, 298–307. neuron disorders: new insights into pathogenic mechanisms. Nat. Rev.
93. Muramatsu, S., Fujimoto, K., Kato, S., Mizukami, H., Asari, S., Genet., 10, 769–782.
Ikeguchi, K., Kawakami, T., Urabe, M., Kume, A., Sato, T. et al. (2010) 113. Ezzi, S.A., Urushitani, M. and Julien, J.P. (2007) Wild-type superoxide
A phase I study of aromatic L-amino acid decarboxylase gene therapy dismutase acquires binding and toxic properties of ALS-linked mutant
for Parkinson’s disease. Mol. Ther., 18, 1731–1735. forms through oxidation. J. Neurochem., 102, 170–178.
94. Feigin, A., Kaplitt, M.G., Tang, C., Lin, T., Mattis, P., Dhawan, V., 114. Bosco, D.A., Morfini, G., Karabacak, N.M., Song, Y., Gros-Louis, F.,
During, M.J. and Eidelberg, D. (2007) Modulation of metabolic brain Pasinelli, P., Goolsby, H., Fontaine, B.A., Lemay, N., McKenna-Yasek,
networks after subthalamic gene therapy for Parkinson’s disease. Proc. D. et al. (2010) Wild-type and mutant SOD1 share an aberrant
Natl Acad. Sci. USA, 104, 19559–19564. conformation and a common pathogenic pathway in ALS. Nat.
95. Lindvall, O. and Kokaia, Z. (2009) Prospects of stem cell therapy for Neurosci., 13, 1369–1403.
replacing dopamine neurons in Parkinson’s disease. Trends Pharmacol. 115. Israelson, A., Arbel, N., Da Cruz, S., Ilieva, H., Yamanaka, K.,
Sci., 30, 260– 267. Shoshan-Barmatz, V. and Cleveland, D.W. (2010) Misfolded mutant
96. Dawson, T.M., Ko, H.S. and Dawson, V.L. (2010) Genetic animal SOD1 directly inhibits VDAC1 conductance in a mouse model of
models of Parkinson’s disease. Neuron, 66, 646– 661. inherited ALS. Neuron, 67, 575–587.
97. Singleton, A.B., Farrer, M., Johnson, J., Singleton, A., Hague, S., 116. Dodge, J.C., Treleaven, C.M., Fidler, J.A., Hester, M., Haidet, A.,
Kachergus, J., Hulihan, M., Peuralinna, T., Dutra, A., Nussbaum, R. Handy, C., Rao, M., Eagle, A., Matthews, J.C., Taksir, T.V. et al. (2010)
et al. (2003) alpha-Synuclein locus triplication causes Parkinson’s AAV4-mediated expression of IGF-1 and VEGF within cellular
disease. Science, 302, 841. components of the ventricular system improves survival outcome in
98. Nemani, V.M., Lu, W., Berge, V., Nakamura, K., Onoa, B., Lee, M.K., familial ALS mice. Mol. Ther., 18, 2075–2084.
Chaudhry, F.A., Nicoll, R.A. and Edwards, R.H. (2010) Increased 117. Dodge, J.C., Haidet, A.M., Yang, W., Passini, M.A., Hester, M., Clarke,
expression of alpha-synuclein reduces neurotransmitter release by J., Roskelley, E.M., Treleaven, C.M., Rizo, L., Martin, H. et al. (2008)
inhibiting synaptic vesicle reclustering after endocytosis. Neuron, 65, Delivery of AAV-IGF-1 to the CNS extends survival in ALS mice
66–79. through modification of aberrant glial cell activity. Mol. Ther., 16,
99. Yamada, M., Mizuno, Y. and Mochizuki, H. (2005) Parkin gene therapy 1056– 1064.
for alpha-synucleinopathy: a rat model of Parkinson’s disease. Hum. 118. Kaspar, B.K., Llado, J., Sherkat, N., Rothstein, J.D. and Gage, F.H.
Gene Ther., 16, 262–270. (2003) Retrograde viral delivery of IGF-1 prolongs survival in a mouse
100. Dominguez, E., Marais, T., Chatauret, N., Benkhelifa-Ziyyat, S., Duque, ALS model. Science, 301, 839–842.
S., Ravassard, P., Carcenac, R., Astord, S., de Moura, A.P., Voit, T. et al. 119. Azzouz, M., Ralph, G.S., Storkebaum, E., Walmsley, L.E.,
(2011) Intravenous scAAV9 delivery of a codon-optimized SMN1 Mitrophanous, K.A., Kingsman, S.M., Carmeliet, P. and Mazarakis, N.D.
sequence rescues SMA mice. Hum. Mol. Genet., 20, 681– 693. (2004) VEGF delivery with retrogradely transported lentivector prolongs
101. Foust, K.D., Wang, X., McGovern, V.L., Braun, L., Bevan, A.K., Haidet, survival in a mouse ALS model. Nature, 429, 413–417.
A.M., Le, T.T., Morales, P.R., Rich, M.M., Burghes, A.H. et al. (2010) 120. Ding, H., Schwarz, D.S., Keene, A., Affarel, B., Fenton, L., Xia, X., Shi,
Rescue of the spinal muscular atrophy phenotype in a mouse model by Y., Zamore, P.D. and Xu, Z. (2003) Selective silencing by RNAi of a
early postnatal delivery of SMN. Nat. Biotechnol., 28, 271– 274. dominant allele that causes amyotrophic lateral sclerosis. Aging Cell, 2,
102. Lorson, C.L., Rindt, H. and Shababi, M. (2010) Spinal muscular 209–217.
atrophy: mechanisms and therapeutic strategies. Hum. Mol. Genet., 19, 121. Ralph, G.S., Radcliffe, P.A., Day, D.M., Carthy, J.M., Leroux, M.A.,
R111–R118. Lee, D.C., Wong, L.F., Bilsland, L.G., Greensmith, L., Kingsman, S.M.
103. Cartegni, L. and Krainer, A.R. (2003) Correction of disease-associated et al. (2005) Silencing mutant SOD1 using RNAi protects against
exon skipping by synthetic exon-specific activators. Nat. Struct. Biol., neurodegeneration and extends survival in an ALS model. Nat. Med., 11,
10, 120–125. 429–433.
104. Williams, J.H., Schray, R.C., Patterson, C.A., Ayitey, S.O., Tallent, M.K. 122. Raoul, C., Abbas-Terki, T., Bensadoun, J.C., Guillot, S., Haase, G.,
and Lutz, G.J. (2009) Oligonucleotide-mediated survival of motor Szulc, J., Henderson, C.E. and Aebischer, P. (2005) Lentiviral-mediated
neuron protein expression in CNS improves phenotype in a mouse model silencing of SOD1 through RNA interference retards disease onset and
of spinal muscular atrophy. J. Neurosci., 29, 7633–7638. progression in a mouse model of ALS. Nat. Med., 11, 423 –428.
105. Hua, Y., Sahashi, K., Hung, G., Rigo, F., Passini, M.A., Bennett, C.F. 123. Xia, X., Zhou, H., Huang, Y. and Xu, Z. (2006) Allele-specific RNAi
and Krainer, A.R. (2010) Antisense correction of SMN2 splicing in the selectively silences mutant SOD1 and achieves significant therapeutic
CNS rescues necrosis in a type III SMA mouse model. Genes Dev., 24, benefit in vivo. Neurobiol. Dis., 23, 578– 586.
1634– 1644. 124. Boillée, S., Yamanaka, K., Lobsiger, C.S., Copeland, N.G., Jenkins,
106. Coady, T.H. and Lorson, C.L. (2010) Trans-splicing-mediated N.A., Kassiotis, G., Kollias, G. and Cleveland, D.W. (2006) Onset and
improvement in a severe mouse model of spinal muscular atrophy. progression in inherited ALS determined by motor neurons and
J. Neurosci., 30, 126–130. microglia. Science, 312, 1389– 1392.
107. Greenway, M.J., Andersen, P.M., Russ, C., Ennis, S., Cashman, S., 125. Urushitani, M., Ezzi, S.A. and Julien, J.P. (2007) Therapeutic effects of
Donaghy, C., Patterson, V., Swingler, R., Kieran, D., Prehn, J. et al. immunization with mutant superoxide dismutase in mice models of
R40 Human Molecular Genetics, 2011, Vol. 20, Review Issue 1

amyotrophic lateral sclerosis. Proc. Natl Acad. Sci. USA, 104, 2495– 145. Sampson, J.H., Archer, G.E., Mitchell, D.A., Heimberger, A.B.,
2500. Herndon, J.E. 2nd, Lally-Goss, D., McGehee-Norman, S., Paolino, A.,
126. Gros-Louis, F., Soucy, G., Lariviere, R. and Julien, J.P. (2010) Reardon, D.A., Friedman, A.H. et al. (2009) An epidermal growth factor
Intracerebroventricular infusion of monoclonal antibody or its derived receptor variant III-targeted vaccine is safe and immunogenic in patients
Fab fragment against misfolded forms of SOD1 mutant delays mortality with glioblastoma multiforme. Mol. Cancer Ther., 8, 2773–2779.
in a mouse model of ALS. J. Neurochem., 113, 1188– 1199. 146. Krisky, D.M., Marconi, P.C., Oligino, T.J., Rouse, R.J., Fink, D.J.,
127. Bleau, A.M., Huse, J.T. and Holland, E.C. (2009) The ABCG2 resistance Cohen, J.B., Watkins, S.C. and Glorioso, J.C. (1998) Development of
network of glioblastoma. Cell Cycle, 8, 2936–2944. herpes simplex virus replication-defective multigene vectors for
128. Plotkin, S.R., Stemmer-Rachamimov, A.O., Barker, F.G. 2nd, Halpin, combination gene therapy applications. Gene Ther., 5, 1517–1530.
C., Padera, T.P., Tyrrell, A., Sorensen, A.G., Jain, R.K. and di Tomaso, 147. Kahlon, K.S., Brown, C., Cooper, L.J., Raubitschek, A., Forman, S.J. and
E. (2009) Hearing improvement after bevacizumab in patients with Jensen, M.C. (2004) Specific recognition and killing of glioblastoma
neurofibromatosis type 2. N. Engl. J. Med., 361, 358– 367. multiforme by interleukin 13-zetakine redirected cytolytic T cells.
129. Lucio-Eterovic, A.K., Piao, Y. and de Groot, J.F. (2009) Mediators of Cancer Res., 64, 9160–9166.
glioblastoma resistance and invasion during antivascular endothelial 148. Maguire, C.A., Meijer, D.H., LeRoy, S.G., Tierney, L.A., Broekman, M.,

Downloaded from https://academic.oup.com/hmg/article-abstract/20/R1/R28/631150 by guest on 15 March 2019


growth factor therapy. Clin. Cancer Res., 15, 4589–4599. Costa, F., Breakefield, X.O., Stemmer-Rachamimov, A. and
130. Määttä, A.M., Samaranayake, H., Pikkarainen, J., Wirth, T. and Sena-Esteves, M. (2008) Preventing growth of brain tumors by creating a
Ylä-Herttuala, S. (2009) Adenovirus mediated herpes simplex zone of resistance. Mol. Ther., 16, 1695– 1702.
virus-thymidine kinase/ganciclovir gene therapy for resectable malignant 149. Jamieson, A.C., Miller, J.C. and Pabo, C.O. (2003) Drug discovery with
glioma. Curr. Gene Ther., 9, 356–367. engineered zinc-finger proteins. Nat. Rev. Drug Discov., 2, 361– 368.
131. Bourbeau, D., Lavoie, G., Nalbantoglu, J. and Massie, B. (2004) Suicide 150. Pâques, F. and Duchateau, P. (2007) Meganucleases and DNA
gene therapy with an adenovirus expressing the fusion gene CD::UPRT double-strand break-induced recombination: perspectives for gene
in human glioblastomas: different sensitivities correlate with p53 status. therapy. Curr. Gene Ther., 7, 49–66.
J. Gene Med., 6, 1320–1332. 151. Simon, P., Cannata, F., Concordet, J.P. and Giovannangeli, C. (2008)
132. Tyminski, E., Leroy, S., Terada, K., Finkelstein, D.M., Hyatt, J.L., Targeting DNA with triplex-forming oligonucleotides to modify gene
Danks, M.K., Potter, P.M., Saeki, Y. and Chiocca, E.A. (2005) Brain sequence. Biochimie, 90, 1109–1116.
tumor oncolysis with replication-conditional herpes simplex virus type 1 152. McLachlan, J., Fernandez, S., Helleday, T. and Bryant, H.E. (2009)
expressing the prodrug-activating genes, CYP2B1 and secreted human Specific targeted gene repair using single-stranded DNA
intestinal carboxylesterase, in combination with cyclophosphamide and oligonucleotides at an endogenous locus in mammalian cells uses
irinotecan. Cancer Res., 65, 6850– 6857. homologous recombination. DNA Repair (Amst.), 8, 1424– 1433.
133. Fulci, G. and Chiocca, E.A. (2007) Suicide gene therapy with an 153. Kazuki, Y., Hoshiya, H., Takiguchi, M., Abe, S., Iida, Y., Osaki, M.,
adenovirus expressing the fusion gene CD::UPRT in human Katoh, M., Hiratsuka, M., Shirayoshi, Y., Hiramatsu, K. et al. (2010)
glioblastomas: different sensitivities correlate with p53 status. Expert Refined human artificial chromosome vectors for gene therapy and
Opin. Biol. Ther., 7, 197– 208. animal transgenesis. Gene Ther., [Epub ahead of print, 18 November].
134. Zemp, F.J., Corredor, J.C., Lun, X., Muruve, D.A. and Forsyth, P.A. 154. Moralli, D., Simpson, K.M., Wade-Martins, R. and Monaco, Z.L. (2006)
(2010) Oncolytic viruses as experimental treatments for malignant A novel human artificial chromosome gene expression system using
gliomas: using a scourge to treat a devil. Cytokine Growth Factor Rev., herpes simplex virus type 1 vectors. EMBO Rep., 7, 911– 918.
21, 103– 117. 155. Bader, A.G., Brown, D. and Winkler, M. (2010) The promise of
135. Goetz, C. and Gromeier, M. (2010) Preparing an oncolytic poliovirus microRNA replacement therapy. Cancer Res., 70, 7027–7030.
recombinant for clinical application against glioblastoma multiforme. 156. Quinti, L., Chopra, V., Rotili, D., Valente, S., Amore, A., Franci, G.,
Cytokine Growth Factor Rev., 21, 197–203. Meade, S., Valenza, M., Altucci, L., Maxwell, M.M. et al. (2010)
136. Lun, X., Chan, J., Zhou, H., Sun, B., Kelly, J.J., Stechishin, O.O., Bell, Evaluation of histone deacetylases as drug targets in Huntington’s
J.C., Parato, K., Hu, K., Vaillant, D. et al. (2010) Efficacy and disease models. Study of HDACs in brain tissues from R6/2 and
safety/toxicity study of recombinant vaccinia virus JX-594 in two CAG140 knock-in HD mouse models and human patients and in a
immunocompetent animal models of glioma. Mol. Ther., 18, neuronal HD cell model. PLoS Curr., 2, RRN1172.
1927–1936. 157. Xie, J., Xie, Q., Zhang, H., Ameres, S.L., Hung, J.H., Su, Q., He, R., Mu,
137. Aboody, K.S., Najbauer, J. and Danks, M.K. (2008) Stem and progenitor X., Seher Ahmed, S., Park, S. et al. (2011) MicroRNA-regulated,
cell-mediated tumor selective gene therapy. Gene Ther., 15, 739–752. systemically delivered rAAV9: a step closer to CNS-restricted transgene
138. Bexell, D., Scheding, S. and Bengzon, J. (2010) Toward brain tumor expression. Mol. Ther., 19, 526– 535.
gene therapy using multipotent mesenchymal stromal cell vectors. Mol. 158. Brown, B.D., Venneri, M.A., Zingale, A., Sergi Sergi, L. and Naldini, L.
Ther., 18, 1067– 1075. (2006) Endogenous microRNA regulation suppresses transgene
139. Frank, R.T., Najbauer, J. and Aboody, K.S. (2010) Concise review: stem expression in hematopoietic lineages and enables stable gene transfer.
cells as an emerging platform for antibody therapy of cancer. Stem Cells, Nat. Med., 12, 585–591.
28, 2084–2087. 159. Brown, B.D., Gentner, B., Cantore, A., Colleoni, S., Amendola, M.,
140. Balyasnikova, I.V., Ferguson, S.D., Sengupta, S., Han, Y. and Lesniak, Zingale, A., Baccarini, A., Lazzari, G., Galli, C. and Naldini, L. (2007)
M.S. (2010) Mesenchymal stem cells modified with a single-chain Endogenous microRNA can be broadly exploited to regulate transgene
antibody against EGFRvIII successfully inhibit the growth of human expression according to tissue, lineage and differentiation state. Nat.
xenograft malignant glioma. PLoS ONE, 5, e9750. Biotechnol., 25, 1457– 1467.
141. Yong, R.L., Shinojima, N., Fueyo, J., Gumin, J., Vecil, G.G., Marini, 160. Mazarakis, N.D., Azzouz, M., Rohll, J.B., Ellard, F.M., Wilkes, F.J.,
F.C., Bogler, O., Andreeff, M. and Lang, F.F. (2009) Human bone Olsen, A.L., Carter, E.E., Barber, R.D., Baban, D.F., Kingsman, S.M.
marrow-derived mesenchymal stem cells for intravascular delivery of et al. (2001) Rabies virus glycoprotein pseudotyping of lentiviral vectors
oncolytic adenovirus Delta24-RGD to human gliomas. Cancer Res., 69, enables retrograde axonal transport and access to the nervous system
8932–8940. after peripheral delivery. Hum. Mol. Genet., 10, 2109– 2121.
142. Herrlinger, U., Woiciechowski, C., Sena-Esteves, M., Aboody, K.S., 161. Rahim, A.A., Wong, A.M., Howe, S.J., Buckley, S.M., Acosta-Saltos,
Jacobs, A.H., Rainov, N.G., Snyder, E.Y. and Breakefield, X.O. (2000) A.D., Elston, K.E., Ward, N.J., Philpott, N.J., Cooper, J.D., Anderson,
Neural precursor cells for delivery of replication-conditional HSV-1 P.N. et al. (2009) Efficient gene delivery to the adult and fetal CNS
vectors to intracerebral gliomas. Mol. Ther., 1, 347– 357. using pseudotyped non-integrating lentiviral vectors. Gene Ther., 16,
143. Shah, K., Bureau, E., Kim, D.E., Yang, K., Tang, Y., Weissleder, R. and 509– 520.
Breakefield, X.O. (2005) Glioma therapy and real-time imaging of neural 162. Maheshri, N., Koerber, J.T., Kaspar, B.K. and Schaffer, D.V. (2006)
precursor cell migration and tumor regression. Ann. Neurol., 57, 34–41. Directed evolution of adeno-associated virus yields enhanced gene
144. Lorico, A., Mercapide, J., Solodushko, V., Alexeyev, M., Fodstad, O. delivery vectors. Nat. Biotechnol., 24, 198– 204.
and Rappa, G. (2008) Primary neural stem/progenitor cells expressing 163. Koerber, J.T., Jang, J.H. and Schaffer, D.V. (2008) DNA shuffling of
endostatin or cytochrome P450 for gene therapy of glioblastoma. Cancer adeno-associated virus yields functionally diverse viral progeny. Mol.
Gene Ther., 15, 605– 615. Ther., 16, 1703–1709.
Human Molecular Genetics, 2011, Vol. 20, Review Issue 1 R41

164. Grimm, D., Lee, J.S., Wang, L., Desai, T., Akache, B., Storm, T.A. and 176. Alper, J. (2009) Geron gets green light for human trial of ES cell-derived
Kay, M.A. (2008) In vitro and in vivo gene therapy vector evolution via product. Nat. Biotechnol., 27, 213– 214.
multispecies interbreeding and retargeting of adeno-associated viruses. 177. Takahashi, K., Tanabe, K., Ohnuki, M., Narita, M., Ichisaka, T.,
J. Virol., 82, 5887–5911. Tomoda, K. and Yamanaka, S. (2007) Induction of pluripotent stem cells
165. Excoffon, K.J., Koerber, J.T., Dickey, D.D., Murtha, M., Keshavjee, S., from adult human fibroblasts by defined factors. Cell, 131, 861 –872.
Kaspar, B.K., Zabner, J. and Schaffer, D.V. (2009) Directed evolution of 178. Yu, J., Vodyanik, M.A., Smuga-Otto, K., Antosiewicz-Bourget, J.,
adeno-associated virus to an infectious respiratory virus. Proc. Natl Frane, J.L., Tian, S., Nie, J., Jonsdottir, G.A., Ruotti, V., Stewart, S. et al.
Acad. Sci. USA, 106, 3865– 3870. (2007) Induced pluripotent stem cell lines derived from human somatic
166. Yang, L., Jiang, J., Drouin, L.M., Agbandje-McKenna, M., Chen, C., cells. Science, 318, 1917– 1920.
Qiao, C., Pu, D., Hu, X., Wang, D.Z., Li, J. et al. (2009) A myocardium 179. Maherali, N. and Hochedlinger, K. (2008) Guidelines and techniques
tropic adeno-associated virus (AAV) evolved by DNA shuffling and in for the generation of induced pluripotent stem cells. Cell Stem Cell, 3,
vivo selection. Proc. Natl Acad. Sci. USA, 106, 3946–3951. 595–605.
167. Gray, S.J., Blake, B.L., Criswell, H.E., Nicolson, S.C., Samulski, R.J. 180. Carey, B.W., Markoulaki, S., Hanna, J., Saha, K., Gao, Q., Mitalipova,
and McCown, T.J. (2010) Directed evolution of a novel adeno-associated M. and Jaenisch, R. (2009) Reprogramming of murine and human

Downloaded from https://academic.oup.com/hmg/article-abstract/20/R1/R28/631150 by guest on 15 March 2019


virus (AAV) vector that crosses the seizure-compromised blood-brain somatic cells using a single polycistronic vector. Proc. Natl Acad. Sci.
barrier (BBB). Mol. Ther., 18, 570–578. USA, 106, 157– 162.
168. Kumar, P., Wu, H., McBride, J.L., Jung, K.E., Kim, M.H., Davidson, 181. Sommer, C.A., Stadtfeld, M., Murphy, G.J., Hochedlinger, K., Kotton,
B.L., Lee, S.K., Shankar, P. and Manjunath, N. (2007) Transvascular D.N. and Mostoslavsky, G. (2009) Induced pluripotent stem cell
delivery of small interfering RNA to the central nervous system. Nature, generation using a single lentiviral stem cell cassette. Stem Cells, 27,
448, 39–43. 543–549.
169. Hajitou, A., Trepel, M., Lilley, C.E., Soghomonyan, S., Alauddin, M.M., 182. Lewitzky, M. and Yamanaka, S. (2007) Reprogramming somatic cells
Marini, F.C., Restel, B.H., Ozawa, M.G., Moya, C.A., Rangel, R. et al.
towards pluripotency by defined factors. Curr. Opin. Biotechnol., 18,
(2006) A hybrid vector for ligand-directed tumor targeting and molecular
467–473.
imaging. Cell, 125, 385–398.
183. Kustikova, O., Fehse, B., Modlich, U., Yang, M., Dullmann, J., Kamino,
170. Staquicini, F.I., Ozawa, M.G., Moya, C.A., Driessen, W.H., Barbu, E.M.,
K., von Neuhoff, N., Schlegelberger, B., Li, Z. and Baum, C. (2005)
Nishimori, H., Soghomonyan, S., Flores, L.G., Liang, X., Paolillo, V.
Clonal dominance of hematopoietic stem cells triggered by retroviral
et al. (2011) Systemic combinatorial peptide selection yields a
non-canonical iron-mimicry mechanism for targeting tumors in a mouse gene marking. Science, 308, 1171–1174.
model of human glioblastoma. J. Clin. Invest., 121, 161 –173. 184. Kane, N.M., Nowrouzi, A., Mukherjee, S., Blundell, M.P., Greig, J.A.,
171. Jiang, L., Rampalli, S., George, D., Press, C., Bremer, E.G., O’Gorman, Lee, W.K., Houslay, M.D., Milligan, G., Mountford, J.C., von Kalle, C.
M.R. and Bohn, M.C. (2004) Tight regulation from a single tet-off rAAV et al. (2010) Lentivirus-mediated reprogramming of somatic cells in the
vector as demonstrated by flow cytometry and quantitative, real-time absence of transgenic transcription factors. Mol. Ther., 18, 2139–2145.
PCR. Gene Ther., 11, 1057–1067. 185. Gutierrez-Aranda, I., Ramos-Mejia, V., Bueno, C., Munoz-Lopez, M.,
172. Candolfi, M., Xiong, W., Yagiz, K., Liu, C., Muhammad, A.K., Puntel, Real, P.J., Macia, A., Sanchez, L., Ligero, G., Garcia-Parez, J.L. and
M., Foulad, D., Zadmehr, A., Ahlzadeh, G.E., Kroeger, K.M. et al. Menendez, P. (2010) Human induced pluripotent stem cells develop
(2010) Gene therapy-mediated delivery of targeted cytotoxins for glioma teratoma more efficiently and faster than human embryonic stem cells
therapeutics. Proc. Natl Acad. Sci. USA, 107, 20021–20026. regardless the site of injection. Stem Cells, 28, 1568–1570.
173. Han, Y., Chang, Q.A., Virag, T., West, N.C., George, D., Castro, M.G. 186. Yamashita, T., Kawai, H., Tian, F., Ohta, Y. and Abe, K. (2010)
and Bohn, M.C. (2010) Lack of humoral immune response to the Tumorigenic development of induced pluripotent stem cells in ischemic
tetracycline (Tet) activator in rats injected intracranially with Tet-off mouse brain. Cell Transplant., [Epub ahead of print, 5 November].
rAAV vectors. Gene Ther., 17, 616–625. 187. Hacein-Bey-Abina, S., Hauer, J., Lim, A., Picard, C., Wang, G.P., Berry,
174. Ye, X., Rivera, V.M., Zoltick, P., Cerasoli, F., Schnell, M.A., Gao, G., C.C., Martinache, C., Rieux-Laucat, F., Latour, S., Belohradsky, B.H.
Hughes, J.V., Gilman, M. and Wilson, J.M. (1999) Regulated delivery of et al. (2010) Efficacy of gene therapy for X-linked severe combined
therapeutic proteins after in vivo somatic cell gene transfer. Science, 283, immunodeficiency. N. Engl. J. Med., 363, 355–364.
88–91. 188. Maguire, A.M., High, K.A., Auricchio, A., Wright, J.F., Pierce, E.A.,
175. Sanftner, L.M., Rivera, V.M., Suzuki, B.M., Feng, L., Berk, L., Zhou, S., Testa, F., Mingozzi, F., Bennicelli, J.L., Ying, G.S., Rossi, S. et al.
Forsayeth, J.R., Clackson, T. and Cunningham, J. (2006) Dimerizer (2009) Age-dependent effects of RPE65 gene therapy for Leber’s
regulation of AADC expression and behavioral response in congenital amaurosis: a phase 1 dose-escalation trial. Lancet, 374,
AAV-transduced 6-OHDA lesioned rats. Mol. Ther., 13, 167–174. 1597– 1605.

Potrebbero piacerti anche