Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2011 Mar 1.
Published in final edited form as: CNS Drugs. 2010 Mar 1;24(3):177–192. doi: 10.2165/11533740-000000000-00000

Gene Therapy in Parkinson’s Disease: Rationale and Current Status

Li Rebekah Feng 1, Kathleen A Maguire-Zeiss 1,*
PMCID: PMC2886503  NIHMSID: NIHMS204668  PMID: 20155994

Abstract

Neurodegenerative diseases pose a unique treatment challenge to clinicians due to the slow progression of disease, the profound neuron loss prior to clinical symptoms and the paucity of early diagnostic biomarkers and restorative therapies. Treatment options are further constrained by the post-mitotic nature of CNS neurons and restricted ability of these cells to regenerate. Lastly, because the blood brain barrier impedes peripheral access to the brain there are inherent limitations with respect to treatment especially protein and peptide-based therapeutics. Due to these intrinsic constraints, researchers are continuing to expand a therapeutic platform based on the delivery of genes engineered for efficient CNS expression. Gene therapeutic approaches were first tested almost 20 years ago and continue to evolve as a viable treatment for CNS neurodegenerative disorders. In this review we consider the current advances in human gene therapy for one common neurodegenerative disorder, Parkinson’s disease (PD).

1. Parkinson’s Disease

The second most common age-related progressive neurodegenerative disorder, PD, first described in 1817 by James Parkinson in his “An Essay on the Shaking Palsy”, affects 1.5 million people in the United States and 4 million people worldwide.(1, 2) Approximately 1–2% of the population over the age of 65 years suffers from PD and this number is expected to grow as the average age of the population increases.(3) The cardinal clinical features include: tremor, cogwheel rigidity, akinesia, bradykinesia and postural instability. However, there is a spectrum among PD patients in terms of clinical manifestations and disease progression. For example, while most patients exhibit the “typical” PD tremor, a rhythmic back-and-forth movement in the range of 4–6 Hz commonly present in the hands or limbs when the patient is at rest (i.e. resting tremor; hands supported against gravity), the severity of these motoric symptoms varies with a subset of PD patients classified as having a tremor-predominant PD phenotype. Although most recognized by the aforementioned debilitating motor dysfunctions, PD patients can also suffer from a variety of non-motor symptoms, including; depression, cognitive impairment, psychosis, hallucinations, compulsive behaviors, REM sleep behavior disorders, excessive daytime somnolence, orthostatic hypotension, gastrointestinal symptoms, constipation, urinary and sexual dysfunction, speech changes, skin problems, pain, difficulty swallowing and chewing.(48) Several of these non-motor complications such as those associated with cognition arise later in disease (Hoehn and Yahr Stages IV–V) while a subset (i.e. gastrointestinal system dysfunction) are prodromal, preceding the motor manifestations by as much as 10 years.(810)

PD is characterized by the loss of substantia nigra pars compacta (SNpc) dopamine neurons, dystrophy of the associated projection fibers to the corpus striatum and diminution of the nigrostriatal neurotransmitter, dopamine. In fact, motoric symptoms are underappreciated until approximately 60% of the SNpc dopamine neurons are lost resulting in a greater than 80% loss of dopamine.(11, 12) The pathological hallmark of PD is presence of intracytoplasmic proteinaceous inclusions called Lewy bodies, which are localized most notably to the few surviving SNpc dopaminergic neurons and were first described by Friederich Lewy in 1912. The function of Lewy bodies in PD pathobiology is not understood. Nonetheless, following immunohistochemical staining with antibodies against two abundant Lewy body proteins, ubiquitin and α-synuclein, the presence of these structures in the remaining SNpc neurons is used as a criteria for diagnosis of PD.(13, 14) The spatial and temporal localization of Lewy bodies has recently been established as a marker for different neuropathological stages of PD. Data from this study demonstrate that Lewy body pathology in PD subjects does not begin in nor is it restricted to the SNpc which corresponds with the observation that non-motor symptoms are also present in this disease.(15) Another feature revealed in PD brains upon autopsy is the presence of activated microglia.(1618) As the immune surveillance cells of the brain, it is not surprising that microglia are upregulated in many neurodegenerative disorders following substantial cell death. However, evidence from both tissue culture and animal models of PD suggest that microglial activation and the consequent increase in proinflammatory molecules may play an important role in the progression of PD, representing another potential therapeutic avenue.(1923)

It has been nearly 200 years since PD was first described and yet the initiating mechanism for sporadic disease has not been fully elucidated. Risk factors including advancing age, exposure to environmental toxicants and drinking well water have all been linked to PD, yet the vast majority of cases are sporadic with no known etiology.(3, 2431) Genetic links to PD have been identified and form the basis for several animal models of this disease.(3235) The first discovered PD-associated gene, SNCA, is localized to chromosome 4 and encodes a 16kDa protein called α-synuclein.(33, 36) Mutations within this gene were linked with an early onset familial form of PD.(34, 37) The importance of α-synuclein in both familial and sporadic PD etiology was further demonstrated following the discovery that duplication or triplication of SNCA was sufficient to cause PD suggesting that overexpression of α-synuclein alone leads to toxicity.(38, 39) Other PD-related genes that have been identified include: parkin, ubiquitin carboxyl-terminal esterase L1 (UCHL1), Leucine-Rich Repeat Kinase 2 (LRRK2), PTEN (phosphatase and tensin homolog)-induced kinase 1 (PINK1) and DJ-1.(35, 4045) Of these genes, LRRK2 mutations are the most frequent and have been linked to both late-onset familial and sporadic PD.(4648) Recent genome-wide association studies revealed several signals associated with PD, one in the gene coding α-synuclein, demonstrating a role for genetic variants in sporadic PD.(49, 50) However, even when taken together, genetic mutations alone do not account for the majority of sporadic PD cases and it has now been generally accepted that a combination of genetic vulnerability and environmental insults lead to sporadic PD.(51) Current clinical diagnosis of PD does not distinguish between initiating factors and usually occurs only after the invariant loss of SNpc dopamine neurons has reached a disease threshold or tipping point. Consequently, existing pharmacotherapies have been focused on enhancing the dysfunctional dopaminergic system, a feature common to all PD patients, rather than halting the disparate initiating events.

2. Rationale for Gene Therapeutic Approaches

All commonly employed PD therapies are focused on the amelioration of symptoms and do not cure disease.(5271) Most pharmaceutical approaches focus on augmentation of the diminished neurotransmitter, dopamine, by either increasing production, extending half-life and/or decreasing metabolism. Dopamine is normally synthesized from the amino acid tyrosine in a series of enzymatic reactions initiated by the conversion of tyrosine to L-dopa by tyrosine hydroxylase and its co-factor tetrahydrobiopterin. L-dopa is then decarboxylated by L-amino acid decarboxylase (AADC) to form dopamine. Dopamine enhancement therapies are most effective when a portion of the nigrostriatal pathway is intact. Consequently, as the number of SNpc dopamine neurons and projection fibers decreases these treatments become less efficacious. The most common therapy for PD is the oral administration of the dopamine precursor, levodopa (L-dopa), usually taken in conjunction with an inhibitor of extracerebral dopa decarboxylase to prevent peripheral metabolism. Introduced more than 40 years ago, this treatment effectively quells the motoric symptoms of akinesia and bradykinesia.(52) However, many patients gradually develop levodopa-induced dyskinesias and motor fluctuations about 5–15 years after the initiation of L-dopa treatment.(72, 73) The severity of the dyskinesias varies between patients and ongoing research efforts are focused on the development of new and more effective anti-dyskinetic medications.(74) A subset of these patients benefit from deep brain stimulation (DBS) which is used in conjunction with L-dopa treatment. Although the exact mechanism of action for DBS is not clear, it is proposed that the generated impulses suppress neural activity, which in the case of subthalamic nucleus DBS (STN-DBS), dampens the PD-increased STN abnormal activity resulting in an overall improvement of tremors, rigidity and akinesias.(54, 57, 60, 62, 65, 6971, 7579) Often DBS-treated patients are able to reduce their L-dopa dose by ~50–60% resulting in a decrease in dyskinesias and report an improved quality of life.(71, 76, 77, 79) Unfortunately, much like L-dopa therapy, the efficacy of DBS declines as PD progresses.

There is no debate regarding the need for novel disease modifying PD therapies. New PD treatments are currently in clinical trials and several are centered on gene therapeutic approaches to either compensate for the loss of dopamine or to protect SNpc dopamine neurons from further degeneration with the overall goal of restoring function (see <clinicaltrials.gov>; Table I; reviewed in (80, 81)). There are several reasons for pursuing a viral vector-mediated gene therapeutic approach in the context of PD; 1) since the PD pathophysiology that subserves the motoric symptoms is largely confined to one brain region, the nigrostriatal pathway, a limited area will require treatment; 2) because of the physically restricted environment of the brain, repeated injections into the nigrostriatum are not desirable, making long-term gene expression following a single treatment appealing; 3) viral vectors are diffusible and theoretically capable of efficient transduction of the striatum; 4) genes have been identified that can either modulate the neuronal phenotype or act as neuroprotective agents; and 5) there is currently no cure for this debilitating disease.

Table I.

Parkinson’s Disease Gene Therapy Clinical Trials

Gene Viral Vector Study Study Phase Sponsor Injection Site ClinicalTrials.gov Identifier Ref
Amino acid decarboxylase rAAV A Study of AAV-hAADC-2 In Subjects With Parkinson’s Disease Phase I Genzyme Striatum NCT00229736 (116, 118, 119, 121, 152156)
Amino acid decarboxylase; Tyrosine hydroxylase; GTP cyclohydrolase I Lentivirus Study of the Safety, Efficacy and Dose Evaluation of ProSavin for the Treatment of Bilateral Idiopathic Parkinson’s Disease Phase I/II Oxford BioMedica Striatum NCT00627588 (107)
Glutamic Acid Decarboxylase (GAD 65/GAD 67) rAAV Safety Study of Subthalamic Nucleus Gene Therapy for Parkinson’s Disease Phase I Neurologix, Inc Subthalamic nucleus NCT00195143 Completed (127129)
Glutamic Acid Decarboxylase (GAD 65/GAD 67) rAAV Study of AAV-GAD Gene Transfer into the Subthalamic Nucleus for Parkinson’s Disease Phase II Neurologix, Inc Subthalamic nucleus NCT00643890 (127129)
Neurturin rAAV Safety of CERE-120 (AAV2-NTN) in Subjects With Idiopathic Parkinson’s Disease Phase I Ceregene Putamen NCT00252850 Completed (144)
Neurturin rAAV Double-Blind, Multicenter, Sham Surgery Controlled Study of CERE-120 in Subjects With Idiopathic Parkinson’s Disease Phase II Ceregene Putamen NCT00400634 Completed (144)

3. Parkinson’s Disease Gene Therapy Platforms

The earliest attempts at PD gene therapy utilized a variety of cell and tissue transplants including fetal and autologous adrenal medullary tissue grafts, xenografts, neurospheres, cell suspension grafts and embryonic stem cells with the overall goal of augmenting dopamine content.(8288) Fetal nerve cell transplantation has been met with some success as a subset of treated patients experienced palliative relief for many years. The effects of long-term fetal implants have recently been evaluated and upon autopsy several subjects displayed PD pathology in the grafted tissue suggesting that the local “disease” environment within the brain brings about “de novo” PD.(8992) However, since there have been only a small number of treated patients available for evaluation it is difficult to draw any definitive conclusions regarding the apparent “transfer of disease”. Regardless these patients demonstrate the potential of cell replacement as an option to stave off disease for several years. Currently one of the biggest obstacles for fetal implant therapy is the acquisition of tissue.

The advent of viral vector technology revolutionized the gene therapy field providing a method for the efficient delivery of genetic material (transgene) without the need for human transplant tissue or cells. Although many different viral platforms have been developed, based on the criterion of safety, stability of gene expression and transduction of specific cellular targets only adeno-associated virus (AAV) and lentivirus vectors are currently in clinical trials for PD (see Table I; for review see (93)). These vectors share the common features of efficient transduction for both dividing and non-dividing cells and long-term transgene expression making them valuable for post-mitotic neuron-targeted therapies.

3.1 Adeno-associated virus vectors

Adeno-associated virus (AAV) is a member of the parvovirus family and a preferred gene therapy vector since this virus is apparently non-pathogenic; in fact, AAV has not been associated with any human disease.(94, 95) Another benefit is that humans, of whom the vast majority have been exposed to AAV, exhibit a low immune response to this virus. Although circulating anti-AAV antibodies have been identified in the human population and innate immune responses to some AAV serotypes have been described, AAV has emerged as the vector platform of choice for gene therapy in humans.(9699) Over 10 recombinant AAV serotypes (rAAV) have been engineered into vectors but rAAV2 is the most frequently employed serotype for gene therapy trials. Additional rAAV serotypes have been developed and tested in animal models that are more efficient at neuronal transduction but these are not yet in clinical trials (Reviewed in (93) and (100103)). The combination of long-term expression, efficient transduction of neurons and diminished proinflammatory and immune responses in humans has thrust rAAV2 to the forefront of PD gene therapy clinical trials (Table 1).

3.2 Lentivirus vectors

Unlike rAAV, which has a restricted transgene size (~4.7 kb), lentivirus vectors can accommodate a larger transgene payload (~8 kb) and are the current vector of choice for multigene PD treatments.(104) Lentiviruses are RNA retroviruses capable of chromosomal integration and stable long-term expression. The most well studied and perhaps widely recognized wild-type lentivirus is the human immunodeficiency virus type I (HIV), the causative agent in acquired immune deficiency syndrome (AIDS). Because of the association of HIV with AIDS the use of lentivirus vectors for gene therapy has raised safety concerns.(105) In order to increase safety, recombinant non-replicating and self-inactivating lentiviruses have been engineered that display tropism for neurons, transduce with high efficiency, and display stable long-term transgene expression.(93, 105, 106) One non-human primate lentiviral vector system based on equine infectious anemia virus (EIAV) with the added capability of self-inactivation has been developed and is presently being used in PD clinical trials (Table I).(107)

4. Parkinson’s Disease Gene Therapy Clinical Trials

Current PD gene therapy clinical trials employ either rAAV2 or lentivirus and are focused on three therapeutic approaches; augmentation of dopamine levels via increased neurotransmitter production, modulation of the neuronal phenotype, and neuroprotection. (106, 108113) One approach focuses on increasing dopamine production via direct delivery of genes involved in neurotransmitter synthesis while a second method is designed to change the neuronal phenotype bypassing the need for dopamine, both approaches should the ameliorate symptoms associated with PD. These therapies are also intended to delay development of end-stage disease, an important accomplishment in a progressive age-related disorder such as PD. On the other hand, delivery of a neurotrophin gene such as glial cell-derived neurotrophic factor (GDNF) or neurturin (NTN), a GDNF-related protein, is projected to slow disease progression by enhancing neuronal survival. However in the long-term it is not known whether neurotrophins, while demonstrated to protect neurons in animal toxicant models of SNpc dopamine neuronal death, will represent a cure for PD. One weakness, which applies to all current PD therapeutic approaches is the lack of a reliable clinical test affording early diagnosis of PD patients prior to the development of motoric symptoms and substantial SNpc dopamine neuron loss.

4.1 Augmentation of dopamine levels

The loss of SNpc dopamine neurons combined with dystrophy of the striatal projection fibers result in a loss of dopamine, which becomes more severe as PD progresses. While oral administration of L-dopa (levodopa) is currently the most effective therapy, ongoing neurodegeneration results in fewer healthy dopamine neurons available to convert this precursor to dopamine. Therefore considerable effort has been placed on the development of therapies aimed at increasing the activity of genes required for dopamine synthesis.(80, 81) These strategies are being tested in Phase I and II clinical trials, as discussed below.

4.1.a. AAV-hAADC-2

Amino acid decarboxylase (AADC) is the enzyme responsible for the decarboxylation of L-dopa to dopamine and enzyme levels have been reported to be lower in PD brains compared with control individuals.(114, 115) These observations led to preclinical studies designed to augment AADC activity. Non-human primate studies demonstrated that gene therapeutic delivery of human AADC via rAAV vector technology (AAV-hAADC-2) restored the ability of the striatum to convert L-dopa into dopamine and mitigated PD-like behavior in toxicant models of disease.(116118) Presumably AAV-hAADC-2 was effective because of the direct transduction of nondegenerating striatal neurons providing a new source of dopamine in addition to increasing the efficiency of the remaining dopamine neurons. This treatment strategy is particularly appealing because it does not require healthy SNpc dopamine neurons, conversion to dopamine is regulated by oral administration of L-dopa and long-term robust gene expression is possible.(116)

The positive preclinical results led to a Phase I clinical trial sponsored by Genzyme Inc. employing the same AAV-hAADC-2 vector to transduce the putamen of ten PD patients (Table I; clinical trial identifier, NCT00229736). The entry criteria included the diagnosis of idiopathic PD with at least two of the four cardinal clinical features, a Hoehn and Yahr Stage III to IV off-medication, intractable motor fluctuations, positive response to L-dopa, a duration of L-dopa therapy ≥ 5 years, age ≤ 75 years and an age at diagnosis ≥ 40 years. Enrolled patients (5 subjects/treatment) received bilateral infusions into the putamen of either low-dose (9 × 1010 vector genomes) or high-dose (3 × 1011 vector genomes) rAAV-AADC-2 in conjunction with orally administered L-dopa to establish safety with a secondary outcome measure aimed at determining the effect of this treatment on clinical status. An initial analysis of 5 treated patients with moderate to advanced PD demonstrated that the rAAV2-hAADC delivery to the striatum was well tolerated and positron emission tomography (PET) using the AADC tracer, [18F]fluoro-L-m-tyrosine (FMT), showed that AADC activity was detectable 6-months post-treatment.(119) Further evidence of safety and tolerability was recently reported for all 10 patients (5 male and 5 female) following 6-months of treatment.(120) Initially 12 patients were screened for this trial but 2 were excluded because they had elevated antibody titers to AAV, of the 10 remaining subjects all reported improvements in the total UPDRS off- and on-states at the 6-month time point. It is also encouraging that 8 of the 10 subjects, all of the high-dose cohort and 3 of the low-dose group, were able to reduce their levodopa dose, an expected outcome if AADC activity was enhanced. AADC activity was monitored with FMT PET and patients in the high dose group showed a greater putaminal uptake compared with the low-dose cohort (75% vs. 30%), which comports with the decrease in the required levodopa dose. Curiously, these subjects showed improvement both off- and on-medication suggesting that augmentation of AADC could increase the conversion efficiency of L-dopa to dopamine for endogenous L-dopa as well as for peripherally administered levodopa. The authors also suggested that sprouting may occur due to glial reaction at the implant site as has been reported for 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP)-treated monkeys undergoing gene transfer and this event may also contribute to the clinical improvement.(120, 121) Future studies will determine whether this is the case in AAV-hAADC-2 treated patients.

Although these studies are encouraging, this is a Phase I clinical trial that was not designed specifically to determine efficacy, another controlled trial is required with a placebo group before benefit directly related to AAV-hAADC-2 can be established. In addition, several adverse effects of treatment most related to the surgical procedure were reported. The most common effects included incisional tenderness, transient headache and in three subjects intracranial hemorrhage, all well described risks following stereotactic craniotomy. In addition, four subjects reported a transient increase in dyskinesias. Patient antibody AAV titers were monitored throughout treatment and increased in 4 of the 6 subjects tested. While no adverse effects were identified related to these increased titers the consequence of circulating AAV antibodies needs to be carefully monitored. Overall this Phase I trial demonstrated that intraputaminal infusion of AAV-hAADC-2 was well tolerated while efficacy needs to be further explored in a controlled Phase II trial. There are two important advantages of this AAV-hAADC-2 strategy; peripherally administered levodopa is used to regulate the amount of dopamine produced (i.e. a “pro-drug” gene transfer) and FMT PET can be used as an unbiased measure of AADC activity.

4.1.b. ProSavin

In a similar approach, Oxford BioMedica in affiliation with the French social security health care system is sponsoring a clinical trial that utilizes a multicistronic lentivirus vector (EIAV-SIN) to transfer 3 transgenes required for the synthesis of dopamine from tyrosine; tyrosine hydroxylase, GTP cyclohydrolase I (required for the synthesis of tetrahydrobiopterin an essential AADC cofactor) and AADC (ProSavin; Lenti-TH-AADC-CH1) into striatal neurons (Table I; clinical trial identifier, NCT00627588). In addition, the tyrosine hydroxylase gene is mutated to prevent the normal negative feedback inhibition of this enzyme by dopamine.(122, 123) Intraputaminal infusions of ProSavin are given in conjunction with peripheral levodopa to modulate the transduced striatal neurons to produce dopamine from both levodopa and endogenous tyrosine making increased amounts of neurotransmitter available for release and binding to post-synaptic dopamine receptors.(107) Similar to the AAV-hAADC-2 treatment, Lenti-TH-AADC-CH1, bypasses the need for healthy SNpc dopamine neurons. In addition, because EIAV-SIN integrates and displays a natural tropism for neurons, long-term expression of the dopamine-modifying genes in striatal neurons can be achieved.

ProSavin therapy has had demonstrated success in animal models of PD and is currently in an open-label PhaseI/II clinical trial (Table I) (106). The initial phase was a dose escalation study to evaluate 2 titers (1X and 2X) in a small cohort of patients (3 subjects/dose). Oxford BioMedica recently announced that both doses were well tolerated, improved motor function and quality of life with no evidence of immunotoxicity or reported adverse effects (16 October 2009). The high dose group showed further motor improvement (34% relative to base line) when evaluated at the 6 months time point using the UPDRS III ‘OFF’ score. Interim results are scheduled for presentation at the European Society of Gene & Cell Therapy Annual Congress in Hanover Germany on November 21–25, 2009. The second stage of the trial is designed to confirm the efficacy of the optimal dose however the preliminary dosage studies lend support for a continued dose escalation study. The Phase II portion is set to enroll 12 patients, 50 to 65 years old, diagnosed with bilateral idiopathic PD for greater than 5 years, with Hoehn and Yahr stage III and IV, motor fluctuations and positive response to dopaminergic therapy.

A potential caveat for this therapy, as well as for the AAV-hAADC-2 gene therapy described above, is that non-dopaminergic striatal neurons are transduced and co-opted to produce and release dopamine. These striatal neurons normally use γ-aminobutyric acid (GABA) as their neurotransmitter and it is unclear what effect producing cytoplasmic dopamine will have on these GABA neurons. For example, cytosolic dopamine is highly reactive and auto-oxidizes to form hydrogen peroxide and semiquinone which can lead to oxidative damage of proteins and lipids, a proposed mechanism for neurodegeneration.(124, 125) Furthermore, it is not clear whether the transduced GABA neurons can behave like dopamine neurons and properly store, release and metabolize this neurotransmitter. In the case of AAV-hAADC-2 therapy these concerns are somewhat reduced since the gene therapy can be controlled by the peripheral administration of levodopa while Lenti-TH-AADC-CH1 therapy will continue to make dopamine from tyrosine in the absence of peripherally supplied levodopa. However, recent work suggests that long-term Lenti-TH-AADC-CH1 therapy may be safe as well as effective since in a non-human primate model of PD, Lenti-TH-AADC-CH1 treatment with an improved vector construct, provided long-term correction of motor deficits (~44 months for 1 animal; 5 additional animals demonstrated improvement up until the 12-month endpoint of the study) without evidence of L-dopa-induced dyskinesias.(126) These preclinical results are encouraging and support further clinical testing of this gene therapeutic approach.

4.2 Modulation of neuronal phenotype

One consequence of the loss of SNpc dopamine neurons is a change in the input to the internal globus pallidus (GPi) and substantia nigra pars reticulata via disinhibition of the subthalamic nucleus (STN). Both STN ablation and deep brain stimulation have been used in patients with advanced PD suggesting that phenotypic modulation of the STN may be beneficial for a subset of patients. An alternative approach to silence the overactive STN includes converting the excitatory neurons to an inhibitory phenotype. In an attempt to modulate the STN neuronal phenotype, one group used rAAV2 to deliver glutamic acid decarboxylase (GAD) to the STN of rats essentially converting glutamatergic neurons to GABA producing cells.(127) This conversion from an excitatory neuron to one capable of inhibitory neurotransmission suppressed the firing activity of the innervated substantia nigra and protected neurons from neurotoxicant-induced degeneration.(127) This neuronal phenotype modulation therapy has moved to Phase I and II clinical trials (Table I; clinical trial identifier, NCT00643890).(108, 127129) The Phase I trial established safety and tolerability. A second randomized, double-blind, placebo controlled trial (Phase II) sponsored by Neurologix, Inc. is underway and enrolling patients that have idiopathic PD (for at least 5 years), UPDRS Part 3 score ≥ 25 in “off” state and demonstrated L-dopa responsiveness for at least 12 months.(129) This is a potentially powerful therapy because it bypasses the degenerating nigrostriatal system but with the caveat that the long-term consequence of “re-programming” glutamatergic/excitatory to GABA/inhibitory neurons is unknown.

4.3 Neuroprotection

Since PD motoric symptoms appear only after substantial loss of SNpc dopamine neurons protection of these cells is an obvious therapeutic goal. Neuroprotection has been achieved in animal models of dopaminergic neuron death following treatment with a prototypical neurotrophic factor, glial cell line-derived neurotrophic factor (GDNF).(130) Although initial clinical trials demonstrated effectiveness, the randomized placebo-controlled double-blinded trial of intraputaminal delivery of the GDNF protein was not successful. Therefore GDNF clinical trials have been halted largely due to the failure of the neurotrophin to reach a large enough target area within the human striatum and more importantly, safety concerns.(55, 131, 132) These concerns arose when neutralizing anti-GDNF antibodies were found in a subset of treated patients suggesting that an immune response might ensue following long-term treatment. In addition, further review of earlier non-human primate data revealed the presence of cerebellar Purkinje cell degeneration suggesting that there was some GDNF peptide leakage outside of the injection site perhaps due to the use of an indwelling catheter.(133) These off-target effects of GDNF and failure to reach the entire striatum speak to the need for gene therapeutic approaches that promote efficient and regulated expression.

Another GDNF family member that supports dopaminergic neurons, neurturin (NTN), was shown to effectively protect dopaminergic neurons in rodent and non-human primate models of PD without the development of neutralizing anti-NTN antibodies or cerebellar degeneration.(55, 134142) In an attempt to enhance temporal and spatial expression of NTN, a viral vector-based platform was chosen to deliver this neurotrophin in human clinical trials rather than direct peptide delivery. Phase I and II trials were completed by Ceregene, Inc. and involved intraputaminal injections of CERE-120, a rAAV2-NTN vector (Table I).(142, 143) Preliminary evidence demonstrated safety and tolerability as well as an improvement in the off-medication motor subscore of the United Parkinson’s Disease Rating Scale (UPDRS), however, other measures of motor function were not significantly improved.(144) Because the viral vector delivery was found to be safe, the Phase II double-blind, multicenter, sham surgery controlled trial was initiated in 58 patients with advanced bilateral idiopathic PD. Disappointingly, in November 2008 Ceregene reported that the Phase II trial did not demonstrate any significant differences in the protocol defined primary endpoint of UPDRS-motor off score at 12 months between patients treated with CERE-120 and control subjects (http://www.ceregene.com/press_112608.asp). However, the neurosurgery and gene therapy were well tolerated with no apparent adverse effects in these advanced-stage patients. In addition, thirty patients were clinically followed in a double-blind fashion for an additional 18 months at which time Ceregene officials reported a modest but statistically significant effect on the UPDRS-motor off score as well as on several secondary measures of motor function (Ceregene, Inc.; http://www.ceregene.com/press_052709.asp). It appears from these studies that over time there was a positive yet minor effect of rAAV2-NTN treatment. The small magnitude of the NTN effect was enigmatic since preclinical studies had indicated robust NTN expression from this vector. However the analysis of brains from two patients enrolled in the Phase II trial that died of causes unrelated to the gene therapy demonstrated NTN expression in the putamen but not within the substantia nigra pars compacta (SNpc). These data support a scenario whereby dystrophic striatal projections are compromised and inefficient in the retrograde transport of NTN to the damaged SNpc dopamine neuron cell bodies. Therefore rAAV2-NTN delivered to the SN where it would be available to directly transduce SN dopamine neurons may be more efficacious. In addition, these clinical trials enrolled patients with advanced PD and presumably significant dopamine neuron loss; neuroprotective therapies may be more effective in earlier stage patients.(81)

5. Conclusion

Recent therapeutic advances have been reported for another neurodegenerative disease, adrenoleukodystrophy (ALD), using hematopoietic stem cells transduced with a lentiviral vector expressing the adenosine triphosphate-binding cassette transporter (ABCD1 gene).(145) These researchers report that ALD progression was halted following treatment. This work supports the concept that neurodegenerative diseases are amenable to gene therapeutic approaches. Although only incremental clinical advancements have been realized thus far for PD gene therapy there is hope that this technology will soon be an important addition to current therapeutic approaches. Gene therapy affords PD clinicians the opportunity to permanently alter dopamine production and neuronal phenotype. Even in the absence of a cure these types of therapies would represent significant therapeutic advancements.

There are critical issues which are currently being addressed that should facilitate the success of PD gene therapy.(133, 142) For example, viral vectors are consistently undergoing improved engineering to enhance delivery, diffusion, and regulated spatial and temporal transgene expression.(104, 146, 147) In addition, effective novel non-invasive imaging techniques to monitor transgene expression and diffusion of viral vectors are under development and will greatly aid the gene therapy field. (148, 149) Furthermore, age-related neurodegenerative diseases such as PD are in need of specific easily accessible and accurate biomarkers that will allow for early diagnosis as well as monitoring of disease progression; these types of studies are currently underway.(112, 133, 150)

It is not clear why preclinical gene therapy studies are more successful than the ensuing clinical trials. Some of the discrepancies could arise from the many different animal models that used for preclinical studies. Researchers use both neurotoxicant and genetic models that do not faithfully recapitulate the progressive degeneration and behavioral aspects of PD. Current work in the field is focused on establishing improved robust animal models. Another important and sometimes overlooked concept is that sporadic PD may represent a syndrome of diseases and therefore therapies may need to be tailored to specific parkinsonian subtypes.(41, 151) Both familial and sporadic PD arise from disparate initiating factors yet the final pathological hallmark is loss of SNpc dopamine neurons. Currently the path between initiation of disease and SNpc dopamine neuron death represents a continuum but with the identification of novel biomarkers clinicians may be able to separate PD patients into subtypes that will benefit from personalized therapies. Furthermore, consistent clinical trial criteria and outcome measurements would allow for reliable analysis of trial data. For example, different PD clinical trials using the same therapeutic molecule (i.e. GDNF) are difficult to interpret because of a lack of standard protocols.(133) Therefore establishing a uniform study design including the same viral vector serotype, injection paradigm, patient selection criteria and outcome measurements would allow for more consistent data interpretation. Finally, PD gene therapy may also benefit from multifaceted approaches, for example gene therapy cocktails such as rAAV2-NTN combined with rAAV2-hAADC and peripheral levodopa may improve both neuronal survival and function better than each treatment alone. In conclusion, as novel targets are identified and the hurdles mentioned above overcome, we anticipate that viral vector-mediated gene therapeutic approaches will have a profound impact on PD progression and outcome.

Acknowledgments

The authors have no conflicts of interest that are directly relevant to the content of this review. This work was funded in part by R01ES014470 (NIEHS) to KMZ.

References

  • 1.Kempster PA, Hurwitz B, Lees AJ. A new look at James Parkinson’s Essay on the Shaking Palsy. Neurology. 2007 Jul 31;69(5):482–5. doi: 10.1212/01.wnl.0000266639.50620.d1. [DOI] [PubMed] [Google Scholar]
  • 2.Parkinson J. An Essay on the Shaking Palsy. London: Sherwood, Neely, and Jones, Paternoster Row; 1817. [Google Scholar]
  • 3.Alves G, Forsaa EB, Pedersen KF, Dreetz Gjerstad M, Larsen JP. Epidemiology of Parkinson’s disease. J Neurol. 2008 Sep;255( Suppl 5):18–32. doi: 10.1007/s00415-008-5004-3. [DOI] [PubMed] [Google Scholar]
  • 4.Dubow JS. Autonomic dysfunction in Parkinson’s disease. Dis Mon. 2007 May;53(5):265–74. doi: 10.1016/j.disamonth.2007.02.004. [DOI] [PubMed] [Google Scholar]
  • 5.Krogh K, Ostergaard K, Sabroe S, Laurberg S. Clinical aspects of bowel symptoms in Parkinson’s disease. Acta Neurol Scand. 2008 Jan;117(1):60–4. doi: 10.1111/j.1600-0404.2007.00900.x. [DOI] [PubMed] [Google Scholar]
  • 6.Poewe W. Dysautonomia and cognitive dysfunction in Parkinson’s disease. Mov Disord. 2007 Sep;22( Suppl 17):S374–8. doi: 10.1002/mds.21681. [DOI] [PubMed] [Google Scholar]
  • 7.Tolosa E, Compta Y, Gaig C. The premotor phase of Parkinson’s disease. Parkinsonism Relat Disord. 2007 Sep;13( Suppl):S2–7. doi: 10.1016/j.parkreldis.2007.06.007. [DOI] [PubMed] [Google Scholar]
  • 8.Grinberg LT, Rueb U, Alho AT, Heinsen H. Brainstem pathology and non-motor symptoms in PD. J Neurol Sci. 2009 Sep 14; doi: 10.1016/j.jns.2009.08.021. [DOI] [PubMed] [Google Scholar]
  • 9.Kaufmann H, Nahm K, Purohit D, Wolfe D. Autonomic failure as the initial presentation of Parkinson disease and dementia with Lewy bodies. Neurology. 2004 Sep 28;63(6):1093–5. doi: 10.1212/01.wnl.0000138500.73671.dc. [DOI] [PubMed] [Google Scholar]
  • 10.Khoo TK, Burn DJ. Non-motor symptoms may herald Parkinson’s disease. Practitioner. 2009 Sep;253(1721):19–24, 2. [PubMed] [Google Scholar]
  • 11.Greffard S, Verny M, Bonnet AM, Beinis JY, Gallinari C, Meaume S, et al. Motor score of the Unified Parkinson Disease Rating Scale as a good predictor of Lewy body-associated neuronal loss in the substantia nigra. Arch Neurol. 2006 Apr;63(4):584–8. doi: 10.1001/archneur.63.4.584. [DOI] [PubMed] [Google Scholar]
  • 12.Toulouse A, Sullivan AM. Progress in Parkinson’s disease-where do we stand? Prog Neurobiol. 2008 Aug;85(4):376–92. doi: 10.1016/j.pneurobio.2008.05.003. [DOI] [PubMed] [Google Scholar]
  • 13.Spillantini MG, Crowther RA, Jakes R, Hasegawa M, Goedert M. alpha-Synuclein in filamentous inclusions of Lewy bodies from Parkinson’s disease and dementia with lewy bodies. Proc Natl Acad Sci U S A. 1998 May 26;95(11):6469–73. doi: 10.1073/pnas.95.11.6469. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Spillantini MG, Schmidt ML, Lee VM, Trojanowski JQ, Jakes R, Goedert M. Alpha-synuclein in Lewy bodies. Nature. 1997 Aug 28;388(6645):839–40. doi: 10.1038/42166. [DOI] [PubMed] [Google Scholar]
  • 15.Braak H, Del Tredici K. Invited Article: Nervous system pathology in sporadic Parkinson disease. Neurology. 2008 May 13;70(20):1916–25. doi: 10.1212/01.wnl.0000312279.49272.9f. [DOI] [PubMed] [Google Scholar]
  • 16.Croisier E, Moran LB, Dexter DT, Pearce RK, Graeber MB. Microglial inflammation in the parkinsonian substantia nigra: relationship to alpha-synuclein deposition. J Neuroinflammation. 2005 Jun 3;2:14. doi: 10.1186/1742-2094-2-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.McGeer PL, Itagaki S, Boyes BE, McGeer EG. Reactive microglia are positive for HLA-DR in the substantia nigra of Parkinson’s and Alzheimer’s disease brains. Neurology. 1988 Aug;38(8):1285–91. doi: 10.1212/wnl.38.8.1285. [DOI] [PubMed] [Google Scholar]
  • 18.Gerhard A, Pavese N, Hotton G, Turkheimer F, Es M, Hammers A, et al. In vivo imaging of microglial activation with [11C](R)-PK11195 PET in idiopathic Parkinson’s disease. Neurobiol Dis. 2006 Feb;21(2):404–12. doi: 10.1016/j.nbd.2005.08.002. [DOI] [PubMed] [Google Scholar]
  • 19.McGeer EG, McGeer PL. The role of anti-inflammatory agents in Parkinson’s disease. CNS Drugs. 2007;21(10):789–97. doi: 10.2165/00023210-200721100-00001. [DOI] [PubMed] [Google Scholar]
  • 20.Su X, Federoff HJ, Maguire-Zeiss KA. Mutant alpha-Synuclein Overexpression Mediates Early Proinflammatory Activity. Neurotox Res. 2009 Oct;16(3):238–54. doi: 10.1007/s12640-009-9053-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Su X, Maguire-Zeiss KA, Giuliano R, Prifti L, Venkatesh K, Federoff HJ. Synuclein activates microglia in a model of Parkinson’s disease. Neurobiol Aging. 2008 Nov;29(11):1690–701. doi: 10.1016/j.neurobiolaging.2007.04.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Lee JK, Tran T, Tansey MG. Neuroinflammation in Parkinson’s Disease. J Neuroimmune Pharmacol. 2009 Oct 10; doi: 10.1007/s11481-009-9176-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Long-Smith CM, Sullivan AM, Nolan YM. The influence of microglia on the pathogenesis of Parkinson’s disease. Prog Neurobiol. 2009 Nov;89(3):277–87. doi: 10.1016/j.pneurobio.2009.08.001. [DOI] [PubMed] [Google Scholar]
  • 24.Gorell JM, Johnson CC, Rybicki BA, Peterson EL, Richardson RJ. The risk of Parkinson’s disease with exposure to pesticides, farming, well water, and rural living. Neurology. 1998 May;50(5):1346–50. doi: 10.1212/wnl.50.5.1346. [DOI] [PubMed] [Google Scholar]
  • 25.Seidler A, Hellenbrand W, Robra BP, Vieregge P, Nischan P, Joerg J, et al. Possible environmental, occupational, and other etiologic factors for Parkinson’s disease: a case-control study in Germany. Neurology. 1996 May;46(5):1275–84. doi: 10.1212/wnl.46.5.1275. [DOI] [PubMed] [Google Scholar]
  • 26.McCormack AL, Thiruchelvam M, Manning-Bog AB, Thiffault C, Langston JW, Cory-Slechta DA, et al. Environmental risk factors and Parkinson’s disease: selective degeneration of nigral dopaminergic neurons caused by the herbicide paraquat. Neurobiol Dis. 2002 Jul;10(2):119–27. doi: 10.1006/nbdi.2002.0507. [DOI] [PubMed] [Google Scholar]
  • 27.Coon S, Stark A, Peterson E, Gloi A, Kortsha G, Pounds J, et al. Whole-body lifetime occupational lead exposure and risk of Parkinson’s disease. Environ Health Perspect. 2006 Dec;114(12):1872–6. doi: 10.1289/ehp.9102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Kamel F, Tanner C, Umbach D, Hoppin J, Alavanja M, Blair A, et al. Pesticide exposure and self-reported Parkinson’s disease in the agricultural health study. Am J Epidemiol. 2007 Feb 15;165(4):364–74. doi: 10.1093/aje/kwk024. [DOI] [PubMed] [Google Scholar]
  • 29.Priyadarshi A, Khuder SA, Schaub EA, Priyadarshi SS. Environmental risk factors and Parkinson’s disease: a metaanalysis. Environ Res. 2001 Jun;86(2):122–7. doi: 10.1006/enrs.2001.4264. [DOI] [PubMed] [Google Scholar]
  • 30.Migliore L, Coppede F. Genetics, environmental factors and the emerging role of epigenetics in neurodegenerative diseases. Mutat Res: Fundam Mol Mech Mutagen. 2008 Oct 31; doi: 10.1016/j.mrfmmm.2008.10.011. [DOI] [PubMed] [Google Scholar]
  • 31.Tanner CM, Ross GW, Jewell SA, Hauser RA, Jankovic J, Factor SA, et al. Occupation and risk of parkinsonism: a multicenter case-control study. Arch Neurol. 2009 Sep;66(9):1106–13. doi: 10.1001/archneurol.2009.195. [DOI] [PubMed] [Google Scholar]
  • 32.Kitada T, Asakawa S, Matsumine H, Hattori N, Shimura H, Minoshima S, et al. Progress in the clinical and molecular genetics of familial parkinsonism. Neurogenetics. 2000;2(4):207–18. doi: 10.1007/s100489900083. [DOI] [PubMed] [Google Scholar]
  • 33.Nussbaum RL, Polymeropoulos MH. Genetics of Parkinson’s disease. Hum Mol Genet. 1997;6(10):1687–91. doi: 10.1093/hmg/6.10.1687. [DOI] [PubMed] [Google Scholar]
  • 34.Polymeropoulos MH. Genetics of Parkinson’s disease. Ann N Y Acad Sci. 2000;920:28–32. doi: 10.1111/j.1749-6632.2000.tb06901.x. [DOI] [PubMed] [Google Scholar]
  • 35.Bras JM, Singleton A. Genetic susceptibility in Parkinson’s disease. Biochim Biophys Acta. 2009 Jul;1792(7):597–603. doi: 10.1016/j.bbadis.2008.11.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Shibasaki Y, Baillie DA, St Clair D, Brookes AJ. High-resolution mapping of SNCA encoding alpha-synuclein, the non-A beta component of Alzheimer’s disease amyloid precursor, to human chromosome 4q21.3-->q22 by fluorescence in situ hybridization. Cytogenet Cell Genet. 1995;71(1):54–5. doi: 10.1159/000134061. [DOI] [PubMed] [Google Scholar]
  • 37.Polymeropoulos MH, Lavedan C, Leroy E, Ide SE, Dehejia A, Dutra A, et al. Mutation in the alpha-synuclein gene identified in families with Parkinson’s disease. Science. 1997;276(5321):2045–7. doi: 10.1126/science.276.5321.2045. [DOI] [PubMed] [Google Scholar]
  • 38.Singleton AB, Farrer M, Johnson J, Singleton A, Hague S, Kachergus J, et al. alpha-Synuclein locus triplication causes Parkinson’s disease. Science. 2003 Oct 31;302(5646):841. doi: 10.1126/science.1090278. [DOI] [PubMed] [Google Scholar]
  • 39.Nishioka K, Hayashi S, Farrer MJ, Singleton AB, Yoshino H, Imai H, et al. Clinical heterogeneity of alpha-synuclein gene duplication in Parkinson’s disease. Ann Neurol. 2006 Feb;59(2):298–309. doi: 10.1002/ana.20753. [DOI] [PubMed] [Google Scholar]
  • 40.Hardy J, Cai H, Cookson MR, Gwinn-Hardy K, Singleton A. Genetics of Parkinson’s disease and parkinsonism. Ann Neurol. 2006 Oct;60(4):389–98. doi: 10.1002/ana.21022. [DOI] [PubMed] [Google Scholar]
  • 41.Hardy J, Lewis P, Revesz T, Lees A, Paisan-Ruiz C. The genetics of Parkinson’s syndromes: a critical review. Curr Opin Genet Dev. 2009 Jun;19(3):254–65. doi: 10.1016/j.gde.2009.03.008. [DOI] [PubMed] [Google Scholar]
  • 42.Liu Z, Meray RK, Grammatopoulos TN, Fredenburg RA, Cookson MR, Liu Y, et al. Membrane-associated farnesylated UCH-L1 promotes alpha-synuclein neurotoxicity and is a therapeutic target for Parkinson’s disease. Proc Natl Acad Sci U S A. 2009 Mar 24;106(12):4635–40. doi: 10.1073/pnas.0806474106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Ragland M, Hutter C, Zabetian C, Edwards K. Association Between the Ubiquitin Carboxyl-Terminal Esterase L1 Gene (UCHL1) S18Y Variant and Parkinson’s Disease: A HuGE Review and Meta-Analysis. Am J Epidemiol. 2009 Oct 28; doi: 10.1093/aje/kwp288. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Sutherland GT, Halliday GM, Silburn PA, Mastaglia FL, Rowe DB, Boyle RS, et al. Do polymorphisms in the familial Parkinsonism genes contribute to risk for sporadic Parkinson’s disease? Mov Disord. 2009 Apr 30;24(6):833–8. doi: 10.1002/mds.22214. [DOI] [PubMed] [Google Scholar]
  • 45.Yasuda T, Nihira T, Ren YR, Cao XQ, Wada K, Setsuie R, et al. Effects of UCH-L1 on alpha-synuclein over-expression mouse model of Parkinson’s disease. J Neurochem. 2009 Feb;108(4):932–44. doi: 10.1111/j.1471-4159.2008.05827.x. [DOI] [PubMed] [Google Scholar]
  • 46.Ferreira JJ, Guedes LC, Rosa MM, Coelho M, van Doeselaar M, Schweiger D, et al. High prevalence of LRRK2 mutations in familial and sporadic Parkinson’s disease in Portugal. Mov Disord. 2007 Jun 15;22(8):1194–201. doi: 10.1002/mds.21525. [DOI] [PubMed] [Google Scholar]
  • 47.Moore DJ. The biology and pathobiology of LRRK2: implications for Parkinson’s disease. Parkinsonism Relat Disord. 2008;14( Suppl 2):S92–8. doi: 10.1016/j.parkreldis.2008.04.010. [DOI] [PubMed] [Google Scholar]
  • 48.Cookson MR, Dauer W, Dawson T, Fon EA, Guo M, Shen J. The roles of kinases in familial Parkinson’s disease. J Neurosci. 2007 Oct 31;27(44):11865–8. doi: 10.1523/JNEUROSCI.3695-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Satake W, Nakabayashi Y, Mizuta I, Hirota Y, Ito C, Kubo M, et al. Genome-wide association study identifies common variants at four loci as genetic risk factors for Parkinson’s disease. Nat Genet. 2009 Dec;41(12):1303–7. doi: 10.1038/ng.485. [DOI] [PubMed] [Google Scholar]
  • 50.Simon-Sanchez J, Schulte C, Bras JM, Sharma M, Gibbs JR, Berg D, et al. Genome-wide association study reveals genetic risk underlying Parkinson’s disease. Nat Genet. 2009 Dec;41(12):1308–12. doi: 10.1038/ng.487. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Maguire-Zeiss KA, Federoff HJ. Convergent pathobiologic model of Parkinson’s disease. Ann N Y Acad Sci. 2003 Jun;991:152–66. doi: 10.1111/j.1749-6632.2003.tb07473.x. [DOI] [PubMed] [Google Scholar]
  • 52.Anden NE, Carlsson A, Kerstell J, Magnusson T, Olsson R, Roos BE, et al. Oral L-dopa treatment of parkinsonism. Acta Med Scand. 1970;187(4):247–55. doi: 10.1111/j.0954-6820.1970.tb02939.x. [DOI] [PubMed] [Google Scholar]
  • 53.Asanuma M, Miyazaki I. Nonsteroidal anti-inflammatory drugs in experimental parkinsonian models and Parkinson’s disease. Curr Pharm Des. 2008;14(14):1428–34. doi: 10.2174/138161208784480153. [DOI] [PubMed] [Google Scholar]
  • 54.Benabid AL. Deep brain stimulation for Parkinson’s disease. Curr Opin Neurobiol. 2003 Dec;13(6):696–706. doi: 10.1016/j.conb.2003.11.001. [DOI] [PubMed] [Google Scholar]
  • 55.Evans JR, Barker RA. Neurotrophic factors as a therapeutic target for Parkinson’s disease. Expert Opin Ther Targets. 2008 Apr;12(4):437–47. doi: 10.1517/14728222.12.4.437. [DOI] [PubMed] [Google Scholar]
  • 56.Factor SA. Current status of symptomatic medical therapy in Parkinson’s disease. Neurotherapeutics. 2008 Apr;5(2):164–80. doi: 10.1016/j.nurt.2007.12.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Guridi J, Obeso JA, Rodriguez-Oroz MC, Lozano AA, Manrique M. L-dopa-induced dyskinesia and stereotactic surgery for Parkinson’s disease. Neurosurgery. 2008 Feb;62(2):311–23. doi: 10.1227/01.neu.0000315998.58022.55. discussion 23–5. [DOI] [PubMed] [Google Scholar]
  • 58.LeWitt PA, Guttman M, Tetrud JW, Tuite PJ, Mori A, Chaikin P, et al. Adenosine A2A receptor antagonist istradefylline (KW-6002) reduces “off” time in Parkinson’s disease: a double-blind, randomized, multicenter clinical trial (6002-US-005) Ann Neurol. 2008 Mar;63(3):295–302. doi: 10.1002/ana.21315. [DOI] [PubMed] [Google Scholar]
  • 59.LeWitt PA, Taylor DC. Protection against Parkinson’s disease progression: clinical experience. Neurotherapeutics. 2008 Apr;5(2):210–25. doi: 10.1016/j.nurt.2008.01.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Limousin P, Martinez-Torres I. Deep brain stimulation for Parkinson’s disease. Neurotherapeutics. 2008 Apr;5(2):309–19. doi: 10.1016/j.nurt.2008.01.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Maguire-Zeiss KA. alpha-Synuclein: A therapeutic target for Parkinson’s disease? Pharmacol Res. 2008 Sep 16; doi: 10.1016/j.phrs.2008.09.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Montgomery EB, Jr, Gale JT. Mechanisms of action of deep brain stimulation(DBS) Neurosci Biobehav Rev. 2008;32(3):388–407. doi: 10.1016/j.neubiorev.2007.06.003. [DOI] [PubMed] [Google Scholar]
  • 63.Prasad KN, Cole WC, Kumar B. Multiple antioxidants in the prevention and treatment of Parkinson’s disease. J Am Coll Nutr. 1999 Oct;18(5):413–23. doi: 10.1080/07315724.1999.10718878. [DOI] [PubMed] [Google Scholar]
  • 64.Ramaker C, van Hilten JJ. Bromocriptine versus levodopa in early Parkinson’s disease. Cochrane Database Syst Rev. 2000;(3):CD002258. doi: 10.1002/14651858.CD002258. [DOI] [PubMed] [Google Scholar]
  • 65.Remple MS, Sarpong Y, Neimat JS. Frontiers in the surgical treatment of Parkinson’s disease. Expert Rev Neurother. 2008 Jun;8(6):897–906. doi: 10.1586/14737175.8.6.897. [DOI] [PubMed] [Google Scholar]
  • 66.Schapira AH, Olanow CW. Rationale for the use of dopamine agonists as neuroprotective agents in Parkinson’s disease. Ann Neurol. 2003;53(Suppl 3):S149–57. doi: 10.1002/ana.10514. discussion S57–9. [DOI] [PubMed] [Google Scholar]
  • 67.Shoulson I. DATATOP: a decade of neuroprotective inquiry. Parkinson Study Group. Deprenyl And Tocopherol Antioxidative Therapy Of Parkinsonism. Ann Neurol. 1998 Sep;44(3 Suppl 1):S160–6. [PubMed] [Google Scholar]
  • 68.Simola N, Morelli M, Pinna A. Adenosine A2A receptor antagonists and Parkinson’s disease: state of the art and future directions. Curr Pharm Des. 2008;14(15):1475–89. doi: 10.2174/138161208784480072. [DOI] [PubMed] [Google Scholar]
  • 69.Wider C, Pollo C, Bloch J, Burkhard PR, Vingerhoets FJ. Long-term outcome of 50 consecutive Parkinson’s disease patients treated with subthalamic deep brain stimulation. Parkinsonism Relat Disord. 2008;14(2):114–9. doi: 10.1016/j.parkreldis.2007.06.012. [DOI] [PubMed] [Google Scholar]
  • 70.Yamada K, Goto S, Hamasaki T, Kuratsu JI. Effect of bilateral subthalamic nucleus stimulation on levodopa-unresponsive axial symptoms in Parkinson’s disease. Acta Neurochir (Wien) 2008 Jan;150(1):15–22. doi: 10.1007/s00701-007-1451-3. discussion. [DOI] [PubMed] [Google Scholar]
  • 71.Yu H, Neimat JS. The treatment of movement disorders by deep brain stimulation. Neurotherapeutics. 2008 Jan;5(1):26–36. doi: 10.1016/j.nurt.2007.10.072. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Fabbrini G, Mouradian MM, Juncos JL, Schlegel J, Mohr E, Chase TN. Motor fluctuations in Parkinson’s disease: central pathophysiological mechanisms, Part I. Ann Neurol. 1988 Sep;24(3):366–71. doi: 10.1002/ana.410240303. [DOI] [PubMed] [Google Scholar]
  • 73.Mouradian MM, Juncos JL, Fabbrini G, Schlegel J, Bartko JJ, Chase TN. Motor fluctuations in Parkinson’s disease: central pathophysiological mechanisms, Part II. Ann Neurol. 1988 Sep;24(3):372–8. doi: 10.1002/ana.410240304. [DOI] [PubMed] [Google Scholar]
  • 74.Del Sorbo F, Albanese A. Levodopa-induced dyskinesias and their management. J Neurol. 2008 Aug;255( Suppl 4):32–41. doi: 10.1007/s00415-008-4006-5. [DOI] [PubMed] [Google Scholar]
  • 75.Herzog J, Pogarell O, Pinsker MO, Kupsch A, Oertel WH, Lindvall O, et al. Deep brain stimulation in Parkinson’s disease following fetal nigral transplantation. Mov Disord. 2008 Jul 15;23(9):1293–6. doi: 10.1002/mds.21768. [DOI] [PubMed] [Google Scholar]
  • 76.Krack P, Batir A, Van Blercom N, Chabardes S, Fraix V, Ardouin C, et al. Five-year follow-up of bilateral stimulation of the subthalamic nucleus in advanced Parkinson’s disease. N Engl J Med. 2003 Nov 13;349(20):1925–34. doi: 10.1056/NEJMoa035275. [DOI] [PubMed] [Google Scholar]
  • 77.Schupbach WM, Chastan N, Welter ML, Houeto JL, Mesnage V, Bonnet AM, et al. Stimulation of the subthalamic nucleus in Parkinson’s disease: a 5 year follow up. J Neurol Neurosurg Psychiatry. 2005 Dec;76(12):1640–4. doi: 10.1136/jnnp.2005.063206. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Deuschl G, Schade-Brittinger C, Krack P, Volkmann J, Schafer H, Botzel K, et al. A randomized trial of deep-brain stimulation for Parkinson’s disease. N Engl J Med. 2006 Aug 31;355(9):896–908. doi: 10.1056/NEJMoa060281. [DOI] [PubMed] [Google Scholar]
  • 79.Schupbach WM, Maltete D, Houeto JL, du Montcel ST, Mallet L, Welter ML, et al. Neurosurgery at an earlier stage of Parkinson disease: a randomized, controlled trial. Neurology. 2007 Jan 23;68(4):267–71. doi: 10.1212/01.wnl.0000250253.03919.fb. [DOI] [PubMed] [Google Scholar]
  • 80.Bjorklund A, Bjorklund T, Kirik D. Gene Therapy for Dopamine Replacement in Parkinson’s Disease. Sci Transl Med. 2009;1(2ps2) doi: 10.1126/scitranslmed.3000350. [DOI] [PubMed] [Google Scholar]
  • 81.Bjorklund T, Kirik D. Scientific rationale for the development of gene therapy strategies for Parkinson’s disease. Biochim Biophys Acta. 2009 Jul;1792(7):703–13. doi: 10.1016/j.bbadis.2009.02.009. [DOI] [PubMed] [Google Scholar]
  • 82.Lindvall O, Backlund EO, Farde L, Sedvall G, Freedman R, Hoffer B, et al. Transplantation in Parkinson’s disease: two cases of adrenal medullary grafts to the putamen. Ann Neurol. 1987 Oct;22(4):457–68. doi: 10.1002/ana.410220403. [DOI] [PubMed] [Google Scholar]
  • 83.Madrazo I, Drucker-Colin R, Diaz V, Martinez-Mata J, Torres C, Becerril JJ. Open microsurgical autograft of adrenal medulla to the right caudate nucleus in two patients with intractable Parkinson’s disease. N Engl J Med. 1987 Apr 2;316(14):831–4. doi: 10.1056/NEJM198704023161402. [DOI] [PubMed] [Google Scholar]
  • 84.Moses D, Drago J, Teper Y, Gantois I, Finkelstein DI, Horne MK. Fetal striatum- and ventral mesencephalon-derived expanded neurospheres rescue dopaminergic neurons in vitro and the nigro-striatal system in vivo. Neuroscience. 2008 June 23;154(2):606–20. doi: 10.1016/j.neuroscience.2008.03.058. [DOI] [PubMed] [Google Scholar]
  • 85.Ostenfeld T, Tai YT, Martin P, Deglon N, Aebischer P, Svendsen CN. Neurospheres modified to produce glial cell line-derived neurotrophic factor increase the survival of transplanted dopamine neurons. J Neurosci Res. 2002 Sep 15;69(6):955–65. doi: 10.1002/jnr.10396. [DOI] [PubMed] [Google Scholar]
  • 86.Sun J, Gao Y, Yang L, Li Z, Lu G, Yew D. Neural-tube-derived neuroepithelial stem cells: a new transplant resource for Parkinson’s disease. Neuroreport. 2007 Apr 16;18(6):543–7. doi: 10.1097/WNR.0b013e3280b07bf4. [DOI] [PubMed] [Google Scholar]
  • 87.Takagi Y, Takahashi J, Saiki H, Morizane A, Hayashi T, Kishi Y, et al. Dopaminergic neurons generated from monkey embryonic stem cells function in a Parkinson primate model. J Clin Invest. 2005 Jan;115(1):102–9. doi: 10.1172/JCI21137. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Takahashi J. Stem cell therapy for Parkinson’s disease. Ernst Schering Res Found Workshop. 2006;(60):229–44. doi: 10.1007/3-540-31437-7_15. [DOI] [PubMed] [Google Scholar]
  • 89.Braak H, Del Tredici K. Assessing fetal nerve cell grafts in Parkinson’s disease. Nat Med. 2008 May;14(5):483–5. doi: 10.1038/nm0508-483. [DOI] [PubMed] [Google Scholar]
  • 90.Kordower JH, Chu Y, Hauser RA, Freeman TB, Olanow CW. Lewy body-like pathology in long-term embryonic nigral transplants in Parkinson’s disease. Nat Med. 2008 May;14(5):504–6. doi: 10.1038/nm1747. [DOI] [PubMed] [Google Scholar]
  • 91.Kordower JH, Chu Y, Hauser RA, Olanow CW, Freeman TB. Transplanted dopaminergic neurons develop PD pathologic changes: a second case report. Mov Disord. 2008 Dec 15;23(16):2303–6. doi: 10.1002/mds.22369. [DOI] [PubMed] [Google Scholar]
  • 92.Li JY, Englund E, Holton JL, Soulet D, Hagell P, Lees AJ, et al. Lewy bodies in grafted neurons in subjects with Parkinson’s disease suggest host-to-graft disease propagation. Nat Med. 2008 May;14(5):501–3. doi: 10.1038/nm1746. [DOI] [PubMed] [Google Scholar]
  • 93.Mandel RJ, Burger C, Snyder RO. Viral vectors for in vivo gene transfer in Parkinson’s disease: properties and clinical grade production. Exp Neurol. 2008 Jan;209(1):58–71. doi: 10.1016/j.expneurol.2007.08.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Bueler H. Adeno-associated viral vectors for gene transfer and gene therapy. Biol Chem. 1999 Jun;380(6):613–22. doi: 10.1515/BC.1999.078. [DOI] [PubMed] [Google Scholar]
  • 95.Samulski RJ, Chang LS, Shenk T. Helper-free stocks of recombinant adeno-associated viruses: normal integration does not require viral gene expression. J Virol. 1989 Sep;63(9):3822–8. doi: 10.1128/jvi.63.9.3822-3828.1989. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Peden CS, Burger C, Muzyczka N, Mandel RJ. Circulating anti-wild-type adeno-associated virus type 2 (AAV2) antibodies inhibit recombinant AAV2 (rAAV2)-mediated, but not rAAV5-mediated, gene transfer in the brain. J Virol. 2004 Jun;78(12):6344–59. doi: 10.1128/JVI.78.12.6344-6359.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Zaiss AK, Muruve DA. Immune responses to adeno-associated virus vectors. Curr Gene Ther. 2005 Jun;5(3):323–31. doi: 10.2174/1566523054065039. [DOI] [PubMed] [Google Scholar]
  • 98.Zaiss AK, Muruve DA. Immunity to adeno-associated virus vectors in animals and humans: a continued challenge. Gene Ther. 2008 Jun;15(11):808–16. doi: 10.1038/gt.2008.54. [DOI] [PubMed] [Google Scholar]
  • 99.Bartlett JS, Samulski RJ, McCown TJ. Selective and rapid uptake of adeno-associated virus type 2 in brain. Hum Gene Ther. 1998 May 20;9(8):1181–6. doi: 10.1089/hum.1998.9.8-1181. [DOI] [PubMed] [Google Scholar]
  • 100.Burger C, Gorbatyuk OS, Velardo MJ, Peden CS, Williams P, Zolotukhin S, et al. Recombinant AAV viral vectors pseudotyped with viral capsids from serotypes 1, 2, and 5 display differential efficiency and cell tropism after delivery to different regions of the central nervous system. Mol Ther. 2004 Aug;10(2):302–17. doi: 10.1016/j.ymthe.2004.05.024. [DOI] [PubMed] [Google Scholar]
  • 101.Mandel RJ, Manfredsson FP, Foust KD, Rising A, Reimsnider S, Nash K, et al. Recombinant adeno-associated viral vectors as therapeutic agents to treat neurological disorders. Mol Ther. 2006 Mar;13(3):463–83. doi: 10.1016/j.ymthe.2005.11.009. [DOI] [PubMed] [Google Scholar]
  • 102.Sondhi D, Hackett NR, Peterson DA, Stratton J, Baad M, Travis KM, et al. Enhanced survival of the LINCL mouse following CLN2 gene transfer using the rh.10 rhesus macaque-derived adeno-associated virus vector. Mol Ther. 2007 Mar;15(3):481–91. doi: 10.1038/sj.mt.6300049. [DOI] [PubMed] [Google Scholar]
  • 103.Reimsnider S, Manfredsson FP, Muzyczka N, Mandel RJ. Time course of transgene expression after intrastriatal pseudotyped rAAV2/1, rAAV2/2, rAAV2/5, and rAAV2/8 transduction in the rat. Mol Ther. 2007 Aug;15(8):1504–11. doi: 10.1038/sj.mt.6300227. [DOI] [PubMed] [Google Scholar]
  • 104.Jakobsson J, Lundberg C. Lentiviral vectors for use in the central nervous system. Mol Ther. 2006 Mar;13(3):484–93. doi: 10.1016/j.ymthe.2005.11.012. [DOI] [PubMed] [Google Scholar]
  • 105.Connolly JB. Lentiviruses in gene therapy clinical research. Gene Ther. 2002 Dec;9(24):1730–4. doi: 10.1038/sj.gt.3301893. [DOI] [PubMed] [Google Scholar]
  • 106.Schambach A, Baum C. Clinical application of lentiviral vectors - concepts and practice. Curr Gene Ther. 2008 Dec;8(6):474–82. doi: 10.2174/156652308786848049. [DOI] [PubMed] [Google Scholar]
  • 107.Azzouz M, Martin-Rendon E, Barber RD, Mitrophanous KA, Carter EE, Rohll JB, et al. Multicistronic lentiviral vector-mediated striatal gene transfer of aromatic L-amino acid decarboxylase, tyrosine hydroxylase, and GTP cyclohydrolase I induces sustained transgene expression, dopamine production, and functional improvement in a rat model of Parkinson’s disease. J Neurosci. 2002 Dec 1;22(23):10302–12. doi: 10.1523/JNEUROSCI.22-23-10302.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Alexander BL, Ali RR, Alton EW, Bainbridge JW, Braun S, Cheng SH, et al. Progress and prospects: gene therapy clinical trials (part 1) Gene Ther. 2007 Oct;14(20):1439–47. doi: 10.1038/sj.gt.3303001. [DOI] [PubMed] [Google Scholar]
  • 109.Fiandaca M, Forsayeth J, Bankiewicz K. Current status of gene therapy trials for Parkinson’s disease. Exp Neurol. 2008 Jan;209(1):51–7. doi: 10.1016/j.expneurol.2007.08.009. [DOI] [PubMed] [Google Scholar]
  • 110.Isacson O, Kordower JH. Future of cell and gene therapies for Parkinson’s disease. Ann Neurol. 2008 Dec;64( Suppl 2):S122–38. doi: 10.1002/ana.21473. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Lewis TB, Standaert DG. Design of clinical trials of gene therapy in Parkinson disease. Exp Neurol. 2008 Jan;209(1):41–7. doi: 10.1016/j.expneurol.2007.08.012. [DOI] [PubMed] [Google Scholar]
  • 112.Maguire-Zeiss KA, Mhyre TR, Federoff HJ. Gazing into the future: Parkinson’s disease gene therapeutics to modify natural history. Exp Neurol. 2008 Jan;209(1):101–13. doi: 10.1016/j.expneurol.2007.09.030. [DOI] [PubMed] [Google Scholar]
  • 113.Mochizuki H, Yasuda T, Mouradian MM. Advances in gene therapy for movement disorders. Neurotherapeutics. 2008 Apr;5(2):260–9. doi: 10.1016/j.nurt.2008.01.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Lloyd K, Hornykiewicz O. Parkinson’s disease: activity of L-dopa decarboxylase in discrete brain regions. Science. 1970 Dec 11;170(963):1212–3. doi: 10.1126/science.170.3963.1212. [DOI] [PubMed] [Google Scholar]
  • 115.Nagatsu T, Yamamoto T, Kato T. A new and highly sensitive voltammetric assay for aromatic L-amino acid decarboxylase activity by high-performance liquid chromatography. Anal Biochem. 1979 Nov 15;100(1):160–5. doi: 10.1016/0003-2697(79)90126-x. [DOI] [PubMed] [Google Scholar]
  • 116.Bankiewicz KS, Forsayeth J, Eberling JL, Sanchez-Pernaute R, Pivirotto P, Bringas J, et al. Long-term clinical improvement in MPTP-lesioned primates after gene therapy with AAV-hAADC. Mol Ther. 2006 Oct;14(4):564–70. doi: 10.1016/j.ymthe.2006.05.005. [DOI] [PubMed] [Google Scholar]
  • 117.Bankiewicz KS, Eberling JL, Kohutnicka M, Jagust W, Pivirotto P, Bringas J, et al. Convection-enhanced delivery of AAV vector in parkinsonian monkeys; in vivo detection of gene expression and restoration of dopaminergic function using pro-drug approach. Exp Neurol. 2000;164(1):2–14. doi: 10.1006/exnr.2000.7408. [DOI] [PubMed] [Google Scholar]
  • 118.Daadi MM, Pivirotto P, Bringas J, Cunningham J, Forsayeth J, Eberling J, et al. Distribution of AAV2-hAADC-transduced cells after 3 years in Parkinsonian monkeys. Neuroreport. 2006 Feb 6;17(2):201–4. doi: 10.1097/01.wnr.0000198952.38563.05. [DOI] [PubMed] [Google Scholar]
  • 119.Eberling JL, Jagust WJ, Christine CW, Starr P, Larson P, Bankiewicz KS, et al. Results from a phase I safety trial of hAADC gene therapy for Parkinson disease. Neurology. 2008 May 20;70(21):1980–3. doi: 10.1212/01.wnl.0000312381.29287.ff. [DOI] [PubMed] [Google Scholar]
  • 120.Christine CW, Starr PA, Larson PS, Eberling JL, Jagust WJ, Hawkins RA, et al. Safety and tolerability of putaminal AADC gene therapy for Parkinson disease. Neurology. 2009 Oct 21; doi: 10.1212/WNL.0b013e3181c29356. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Sanftner LM, Sommer JM, Suzuki BM, Smith PH, Vijay S, Vargas JA, et al. AAV2-mediated gene delivery to monkey putamen: evaluation of an infusion device and delivery parameters. Exp Neurol. 2005 Aug;194(2):476–83. doi: 10.1016/j.expneurol.2005.03.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Kumer SC, Vrana KE. Intricate regulation of tyrosine hydroxylase activity and gene expression. J Neurochem. 1996 Aug;67(2):443–62. doi: 10.1046/j.1471-4159.1996.67020443.x. [DOI] [PubMed] [Google Scholar]
  • 123.Wachtel SR, Bencsics C, Kang UJ. Role of aromatic L-amino acid decarboxylase for dopamine replacement by genetically modified fibroblasts in a rat model of Parkinson’s disease. J Neurochem. 1997 Nov;69(5):2055–63. doi: 10.1046/j.1471-4159.1997.69052055.x. [DOI] [PubMed] [Google Scholar]
  • 124.Cheng N, Maeda T, Kume T, Kaneko S, Kochiyama H, Akaike A, et al. Differential neurotoxicity induced by L-DOPA and dopamine in cultured striatal neurons. Brain Res. 1996 Dec 16;743(1–2):278–83. doi: 10.1016/s0006-8993(96)01056-6. [DOI] [PubMed] [Google Scholar]
  • 125.Graham DG, Tiffany SM, Bell WR, Jr, Gutknecht WF. Autoxidation versus covalent binding of quinones as the mechanism of toxicity of dopamine, 6-hydroxydopamine, and related compounds toward C1300 neuroblastoma cells in vitro. Mol Pharmacol. 1978 Jul;14(4):644–53. [PubMed] [Google Scholar]
  • 126.Jarrya B, Boulet S, Ralph GS, Jan C, Bonvento G, Azzouz M, et al. Dopamine Gene Therapy for Parkinson’s Disease in a Nonhuman Primate Without Associated Dyskinesia. Sci Transl Med. 2009;1(2ra4) doi: 10.1126/scitranslmed.3000130. [DOI] [PubMed] [Google Scholar]
  • 127.Luo J, Kaplitt MG, Fitzsimons HL, Zuzga DS, Liu Y, Oshinsky ML, et al. Subthalamic GAD gene therapy in a Parkinson’s disease rat model. Science. 2002 Oct 11;298(5592):425–9. doi: 10.1126/science.1074549. [DOI] [PubMed] [Google Scholar]
  • 128.During MJ, Kaplitt MG, Stern MB, Eidelberg D. Subthalamic GAD gene transfer in Parkinson disease patients who are candidates for deep brain stimulation. Hum Gene Ther. 2001 Aug 10;12(12):1589–91. [PubMed] [Google Scholar]
  • 129.Kaplitt MG, Feigin A, Tang C, Fitzsimons HL, Mattis P, Lawlor PA, et al. Safety and tolerability of gene therapy with an adeno-associated virus (AAV) borne GAD gene for Parkinson’s disease: an open label, phase I trial. Lancet. 2007 Jun 23;369(9579):2097–105. doi: 10.1016/S0140-6736(07)60982-9. [DOI] [PubMed] [Google Scholar]
  • 130.Choi-Lundberg DL, Lin Q, Chang YN, Chiang YL, Hay CM, Mohajeri H, et al. Dopaminergic neurons protected from degeneration by GDNF gene therapy. Science. 1997 Feb 7;275(5301):838–41. doi: 10.1126/science.275.5301.838. [DOI] [PubMed] [Google Scholar]
  • 131.Nutt JG, Burchiel KJ, Comella CL, Jankovic J, Lang AE, Laws ER, Jr, et al. Randomized, double-blind trial of glial cell line-derived neurotrophic factor (GDNF) in PD. Neurology. 2003 Jan 14;60(1):69–73. doi: 10.1212/wnl.60.1.69. [DOI] [PubMed] [Google Scholar]
  • 132.Barker RA. Continuing trials of GDNF in Parkinson’s disease. Lancet Neurol. 2006 Apr;5(4):285–6. doi: 10.1016/S1474-4422(06)70386-6. [DOI] [PubMed] [Google Scholar]
  • 133.Sherer TB, Fiske BK, Svendsen CN, Lang AE, Langston JW. Crossroads in GDNF therapy for Parkinson’s disease. Mov Disord. 2006 Feb;21(2):136–41. doi: 10.1002/mds.20861. [DOI] [PubMed] [Google Scholar]
  • 134.Bespalov MM, Saarma M. GDNF family receptor complexes are emerging drug targets. Trends Pharmacol Sci. 2007 Feb;28(2):68–74. doi: 10.1016/j.tips.2006.12.005. [DOI] [PubMed] [Google Scholar]
  • 135.Gasmi M, Brandon EP, Herzog CD, Wilson A, Bishop KM, Hofer EK, et al. AAV2-mediated delivery of human neurturin to the rat nigrostriatal system: long-term efficacy and tolerability of CERE-120 for Parkinson’s disease. Neurobiol Dis. 2007 Jul;27(1):67–76. doi: 10.1016/j.nbd.2007.04.003. [DOI] [PubMed] [Google Scholar]
  • 136.Gasmi M, Herzog CD, Brandon EP, Cunningham JJ, Ramirez GA, Ketchum ET, et al. Striatal delivery of neurturin by CERE-120, an AAV2 vector for the treatment of dopaminergic neuron degeneration in Parkinson’s disease. Mol Ther. 2007 Jan;15(1):62–8. doi: 10.1038/sj.mt.6300010. [DOI] [PubMed] [Google Scholar]
  • 137.Gottwald MD, Aminoff MJ. New frontiers in the pharmacological management of Parkinson’s disease. Drugs Today (Barc) 2008 Jul;44(7):531–45. doi: 10.1358/dot.2008.44.7.1217105. [DOI] [PubMed] [Google Scholar]
  • 138.Herzog CD, Dass B, Gasmi M, Bakay R, Stansell JE, Tuszynski M, et al. Transgene expression, bioactivity, and safety of CERE-120 (AAV2-neurturin) following delivery to the monkey striatum. Mol Ther. 2008 Oct;16(10):1737–44. doi: 10.1038/mt.2008.170. [DOI] [PubMed] [Google Scholar]
  • 139.Herzog CD, Brown L, Gammon D, Kruegel B, Lin R, Wilson A, et al. Expression, bioactivity, and safety 1 year after adeno-associated viral vector type 2-mediated delivery of neurturin to the monkey nigrostriatal system support cere-120 for Parkinson’s disease. Neurosurgery. 2009 Apr;64(4):602–12. doi: 10.1227/01.NEU.0000340682.06068.01. discussion 12–3. [DOI] [PubMed] [Google Scholar]
  • 140.Peterson AL, Nutt JG. Treatment of Parkinson’s disease with trophic factors. Neurotherapeutics. 2008 Apr;5(2):270–80. doi: 10.1016/j.nurt.2008.02.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Rosenblad C, Kirik D, Devaux B, Moffat B, Phillips HS, Bjorklund A. Protection and regeneration of nigral dopaminergic neurons by neurturin or GDNF in a partial lesion model of Parkinson’s disease after administration into the striatum or the lateral ventricle. Eur J Neurosci. 1999 May;11(5):1554–66. doi: 10.1046/j.1460-9568.1999.00566.x. [DOI] [PubMed] [Google Scholar]
  • 142.Bartus RT, Herzog CD, Bishop K, Ostrove JM, Tuszynski M, Kordower JH, et al. Issues regarding gene therapy products for Parkinson’s disease: the development of CERE-120 (AAV-NTN) as one reference point. Parkinsonism Relat Disord. 2007;13( Suppl 3):S469–77. doi: 10.1016/S1353-8020(08)70052-X. [DOI] [PubMed] [Google Scholar]
  • 143.Kotzbauer PT, Lampe PA, Heuckeroth RO, Golden JP, Creedon DJ, Johnson EM, Jr, et al. Neurturin, a relative of glial-cell-line-derived neurotrophic factor. Nature. 1996 Dec 5;384(6608):467–70. doi: 10.1038/384467a0. [DOI] [PubMed] [Google Scholar]
  • 144.Marks WJ, Jr, Ostrem JL, Verhagen L, Starr PA, Larson PS, Bakay RA, et al. Safety and tolerability of intraputaminal delivery of CERE-120 (adeno-associated virus serotype 2-neurturin) to patients with idiopathic Parkinson’s disease: an open-label, phase I trial. Lancet Neurol. 2008 May;7(5):400–8. doi: 10.1016/S1474-4422(08)70065-6. [DOI] [PubMed] [Google Scholar]
  • 145.Cartier N, Hacein-Bey-Abina S, Bartholomae CC, Veres G, Schmidt M, Kutschera I, et al. Hematopoietic stem cell gene therapy with a lentiviral vector in X-linked adrenoleukodystrophy. Science. 2009 Nov 6;326(5954):818–23. doi: 10.1126/science.1171242. [DOI] [PubMed] [Google Scholar]
  • 146.Haberman RP, McCown TJ. Regulation of gene expression in adeno-associated virus vectors in the brain. Methods. 2002 Oct;28(2):219–26. doi: 10.1016/s1046-2023(02)00226-8. [DOI] [PubMed] [Google Scholar]
  • 147.Maguire-Zeiss KA, Federoff HJ. Safety of viral vectors for neurological gene therapies. Curr Opin Mol Ther. 2004;6(5):473–81. [PubMed] [Google Scholar]
  • 148.Asokan A, Johnson JS, Li C, Samulski RJ. Bioluminescent virion shells: new tools for quantitation of AAV vector dynamics in cells and live animals. Gene Ther. 2008 Dec;15(24):1618–22. doi: 10.1038/gt.2008.127. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Fiandaca MS, Varenika V, Eberling J, McKnight T, Bringas J, Pivirotto P, et al. Real-time MR imaging of adeno-associated viral vector delivery to the primate brain. Neuroimage. 2008 Nov 27; doi: 10.1016/j.neuroimage.2008.11.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Ravina B, Tanner C, Dieuliis D, Eberly S, Flagg E, Galpern WR, et al. A longitudinal program for biomarker development in Parkinson’s disease: a feasibility study. Mov Disord. 2009 Oct 30;24(14):2081–90. doi: 10.1002/mds.22690. [DOI] [PubMed] [Google Scholar]
  • 151.Klein C, Schlossmacher MG. Parkinson disease, 10 years after its genetic revolution: multiple clues to a complex disorder. Neurology. 2007 Nov 27;69(22):2093–104. doi: 10.1212/01.wnl.0000271880.27321.a7. [DOI] [PubMed] [Google Scholar]
  • 152.Bankiewicz KS, Daadi M, Pivirotto P, Bringas J, Sanftner L, Cunningham J, et al. Focal striatal dopamine may potentiate dyskinesias in parkinsonian monkeys. Exp Neurol. 2006 Feb;197(2):363–72. doi: 10.1016/j.expneurol.2005.10.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Cunningham J, Pivirotto P, Bringas J, Suzuki B, Vijay S, Sanftner L, et al. Biodistribution of adeno-associated virus type-2 in nonhuman primates after convection-enhanced delivery to brain. Mol Ther. 2008 Jul;16(7):1267–75. doi: 10.1038/mt.2008.111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Eberling JL, Bankiewicz KS, O’Neil JP, Jagust WJ. PET 6-[F]fluoro-L-m-tyrosine Studies of Dopaminergic Function in Human and Nonhuman Primates. Front Hum Neurosci. 2007;1:9. doi: 10.3389/neuro.09.009.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Forsayeth JR, Eberling JL, Sanftner LM, Zhen Z, Pivirotto P, Bringas J, et al. A dose-ranging study of AAV-hAADC therapy in Parkinsonian monkeys. Mol Ther. 2006 Oct;14(4):571–7. doi: 10.1016/j.ymthe.2006.04.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Hadaczek P, Kohutnicka M, Krauze MT, Bringas J, Pivirotto P, Cunningham J, et al. Convection-enhanced delivery of adeno-associated virus type 2 (AAV2) into the striatum and transport of AAV2 within monkey brain. Hum Gene Ther. 2006 Mar;17(3):291–302. doi: 10.1089/hum.2006.17.291. [DOI] [PubMed] [Google Scholar]

RESOURCES