Personal Use Only Not for Distribution

0 downloads 0 Views 6MB Size Report
13621-9. [111] Lewis J, Melrose H, Bumcrot D, et al. .... [143] Dansithong W, Paul S, Scoles DR, et al. Generation of ..... study. Neurosci Lett 2011; 492(2): 94-8.
Send Orders for Reprints to [email protected] Current Gene Therapy, 2018, 18, 206-224

REVIEW ARTICLE ISSN: 1566-5232 eISSN: 1875-5631

Cellular, Molecular and Non-pharmacological Therapeutic Advances for the Treatment of Parkinson's Disease: Separating Hope from Hype

Impact Factor:

1.94

BENTHAM SCIENCE

Mehdi Ghamgosha1, Ali Mohammad Latifi1, Gholam Hossein Meftahi2 and Alireza Mohammadi2,* 1

Applied Biotechnology Research Center, Baqiyatallah University of Medical Sciences, Tehran, Iran; 2Neuroscience Research Center, Baqiyatallah University of Medical Sciences, Tehran, Iran

ARTICLE HISTORY Received: April 10, 2018 Revised: August 24, 2018 Accepted: September 05, 2018 DOI: 10.2174/1566523218666180910163401

Abstract: Parkinson’s Disease (PD) is a frustrating condition characterized by motor and nonmotor deficits majorly caused by the loss  of dopaminergic cells in the Substantia Nigra pars compacta (SNc) and destruction of the nigrostriatal pathway. Despite the very respectable advances in cutting-edge approaches for the treatment of PD, there exist numerous challenges that have incapacitated the definitive treatment of this disease. This review emphasized the development of various non-pharmaceutical therapeutic approaches and mainly highlighted the cutting-edge treatments for PD including gene- and stem cell-based therapies, targeted delivery of neurotrophic factors, and brain stimulation techniques such as Transcranial Magnetic Stimulation (TMS), transcranial Direct Current Stimulation (tDCS), and Deep Brain Stimulation (DBS). The review covered various gene therapy strategies including Adeno-Associated Virus-Glutamic Acid Decarboxylase (AAV-GAD), AAV–Aromatic L-Amino Acid Decarboxylase (AAV-AADC), Lenti-AADC/Tyrosine Hydroxylase/Guanosine TriphosphateCyclohydrolase I (Lenti-AADC/TH/GTP-CH1), AAV–Neurturin (AAV-NRTN), α -Synuclein silencing, and PRKN gene delivery. Also, the advantages, disadvantages, and the results of trials of these methods were discussed. Finally, reasons for the failure of PD treatment were described, with the hopes separated from hypes.

Pe N rs ot on fo al rD U is se tri O bu n tio ly n

Current Gene Therapy

206

Keywords: PRKN gene delivery, α-Synuclein silencing, tDCS, DBS, Cell therapy, AAV-GAD. 1. INTRODUCTION

PD is a frustrating condition characterized by motor and nonmotor deficits majorly caused by the loss  of dopaminergic cells in the SNc and destruction of the nigrostriatal pathway [1, 2]. The nigrostriatal pathway links the SNc with the dorsal striatum (the putamen and caudate nucleus), and the devastation of neurons in this pathway lead to cardinal symptoms of PD, including tremor, rigidity, postural instability, and akinesia. Though brain imaging and electroencephalographic studies are applied in the early diagnosis of neurological disorders [3, 4], PD symptoms become evident when about 60% of dopaminergic neurons have been destroyed [1]. Although the exact cellular and molecular mechanisms leading to the degeneration of dopaminergic cells in PD are unclear, the identification of several mutations in SNCA (PARK1 or PARK4), PRKN (PARK2 or parkin RBR E3 ubiquitin protein ligase), GAK (cyclin G-associated kinase), GBA (β-glucocerebrosidase), LRRK2 (Leucine-rich repeat kinase 2 or PARK8), DJ-1 (PARK7), PARK16, NAT2 (Nacetyltransferase 2), PINK1 (PTEN-induced putative kinase 1 or PARK6), HLA-DRA and ATP13A2 (PARK9) genes *Address correspondence to this author at the Neuroscience Research Center, Baqiyatallah University of Medical Sciences, P.O. Box: 19395-6558, Postal Code: 1413643561, Tehran, Iran; Tel: +98 (21) 26127286; E-mail:  [email protected] 1875-5631/18 $58.00+.00

have made an inimitable glance into the mechanisms responsible for the etiology of PD [5-10]. The SNCA gene is recognized as the main causative gene responsible for producing a protein with 140 amino acids, α -synuclein (α-syn). It has been previously stated that α-syn may be a chief element of Lewy Bodies (LBs), a well-known neuropathological biomarker of PD. LBs are usually diffused through the substantia nigra, hypothalamus, locus ceruleus, and cerebral cortex [11]. Several studies have shown that mitochondrial dysfunction may occur due to loss of function of PRKN, PINK1, LRRK2, and DJ-1 [12-16]. Moreover, it has been revealed that mitochondrial dysfunction, lysosomal degradation, overproduction of Reactive Oxygen Species (ROS), reduced antioxidant molecules (e.g., DJ-1) and oxidatively modified dopamine (DA-Ox) accelerate the accumulation of α-syn in the oligomers, fibrils and eventually LBs (Fig. 1) [17, 18]. Current pharmacological therapies for PD (levodopa, Catechol-O-Methyltransferase (COMT) inhibitors, dopamine receptor agonists, and monoamine oxidase B inhibitors) led to increased activity of the remaining dopaminergic neurons. The useful effects of levodopa as the gold standard of these drugs decrease by way of the remaining dopaminergic cells losing their ability over time [19]. Therefore, effective treatment for PD has mostly failed because of these limitations and the late diagnosis of the disease. In  this  review, the de-

© 2018 Bentham Science Publishers

Therapeutic Advances in the Treatment of PD

Current Gene Therapy, 2018, Vol. 18, No. 4

lysosome

Mitochondrial dysfunction LRRK2

PINK1

Gcase ALP

PRKN

LRRK2 PINK1 PRKN DJ-1

DJ-1

Ab oligomers

207

DJ-1

ROS

DJ-1 Hyperphosphorylated tau Ub

Ub

Ub

Ub

P

P

P

Ub

DA

UPS

Lewy neurites

Ox-

Pe N rs ot on fo al rD U is se tri O bu n tio ly n

Ub

P P

Ub Ub Ub

Lewy body

a-synuclein Ub Ub Ub Ub Ub Ub

a-Synuclein oligomers

Healthy substantia nigra

Lewy body

Diminished substantia nigra

Fig. (1). The pathophysiological mechanisms of Lewy body formation in PD. Loss-of-function or mutations in PRKN, PINK1, and DJ-1 due to mitochondrial dysfunction and induction of autophagy. LRRK2 is involved in chaperone-mediated autophagy and phosphorylation of Microtubuleassociated Protein Tau (MAPT) and α -syn. The ubiquitin-proteasome system eliminates misfolded and damaged proteins from the cytosol, endoplasmic reticulum, and nucleus. Auto-oxidation of dopamine (Ox-DA), lysosomal degradation, Reactive Oxygen Species (ROS) overproduction, and decreased redox-sensitive chaperone molecules like DJ-1, accelerates the accumulation of α-syn in the oligomers, fibrils, and eventually LBs. Aβ and tau considerably induce the assembly of α-syn into fibrils.

velopment of various non-pharmaceutical therapeutic approaches will be emphasized and the cutting-edge treatments for PD including gene- and stem cell-based therapies, targeted delivery of neurotrophic factors, and brain stimulation techniques such as transcranial direct current stimulation (tDCS), Transcranial Magnetic Stimulation (TMS), and Deep Brain Stimulators (DBS) highlighted. Also, the advantages and disadvantages of these methods will be discussed, and the hopes separated from hypes. 2. GENE THERAPY

Gene therapy for PD has obvious potential benefits as it can increase the accessibility of dopamine synthesis precursors and can be used to retain and/or repair dopaminergic neurons as well [20-23]. Gene therapy can correct a specific genetic defect by increasing, decreasing or silencing the expression of target genes, or induction of endogenous production of a therapeutic protein. Transferring genes encoding therapeutic proteins to targeted regions of the brain like SNc or striatum may be a hopeful moment for the improvement of motor and non-motor complications of PD. It has been illustrated that the loss of dopaminergic cells is delayed by neurturin (NRTN) or Glial cell line-derived Neurotrophic

Factor (GDNF) recombinant proteins delivery to the SNc or striatum [20, 24-29]. On the other hand, the polarity and large size of these proteins lead to protein misfolding and consequently their inactivity. Other disadvantages include the inability to easily cross the Blood-Brain Barrier (BBB) and, in some cases, significant side effects that reduce their therapeutic value [30-33]. Therefore, the peripherally administered vectors are not able to cross the BBB and necessitate direct injecting into the brain. Although gene therapy has some shortcomings and disadvantages, the improvements indicate its ability to improve dopaminergic cells and help to resolve motor and non-motor problems in Parkinson’s. Hence, gene delivery may be an efficient approach to convey trophic factors to the SNc or striatum of patients with PD. 2.1. Viral Vectors Recombinant viruses as the most proficient vehicles to realize long-term constant gene expression has been extensively utilized as vectors for the production of therapeutic proteins. The most commonly used vectors for gene delivery to the brain are based on lentivirus (for example HIV-1), herpes simplex virus (like type 1 or HSV-1), adenovirus,

208 Current Gene Therapy, 2018, Vol. 18, No. 4

Adeno-Associated Virus (AAV), retrovirus, SV40, equine infectious anemia virus, and feline immunodeficiency virus. AAV and lentivirus vectors are used as the non- or low-toxic eminent vectors for gene delivery to the brain [34-38], however, it has been shown that AAV vectors are capable of exciting T lymphocyte cytotoxic response [39].

[51, 57]. These findings were completed in 13 adult MPTPinduced hemiparkinsonian rhesus monkeys, as [18F] fluorodeoxyglucose positron emission tomography (FDG-PET) investigation showed that the STN AAV-GAD treatment resulted in an increased glucose metabolism of motor cortex and significant clinical rating improvement [50]. Furthermore, behavioural but not neuroprotective effect of STN AAV-GAD65 treatment have been replicated in a 6-OHDAlesioned rat model of PD [42, 57]. In contrast, AAV genomes were detected in STN but not in the blood or any other tissues of STN rAAV2-GAD-treated rat model of PD [42]. The first clinical trial AAV–GAD gene therapy in patients with PD was conducted in an open-label nonrandomized phase I trial which dates back to 2007. Kaplitt and colleagues injected unilateral STN AAV2-GAD65/67 into 12 (11 men and 1 woman) PD patients and noteworthy improvements were observed 3 to 12 months post-surgery in the motor Unified Parkinson’s Disease Rating Scale (UPDRS) scores. From their findings, AAV2-GAD65/67 gene therapy of the STN was harmless and well tolerated by patients with advanced PD [58]. Subsequently, Feigin and coworkers (2007) confirmed these findings using FDG-PET and stated that unilateral STN AAV-GAD gene therapy leads to a reduction of metabolism in the pallidum and thalamus of the operated side [59]. It has also been revealed that the improved motor UPDRS scores are associated with increases in the premotor cortex metabolism of the same hemispheres [60]. A randomized, double-blind, phase II trial of bilateral STN AAV2-GAD65/67 in 45 patients with progressive levodopa-responsive PD, with significantly low motor UPDRS scores, were compared with controls. The results of 6 months monitoring revealed a considerable improvement of STN AAV2-GAD group in motor UPDRS scores than that of the controls. The safety and efficacy of STN AAV2GAD were listed as the strengths of this study, and nausea, headache, and a bowel obstruction as its side effects; however, it was stated that the last case was not attributed to the intervention or surgical procedure [56]. In a newly reported bilateral STN AAV2-GAD randomized and double-blind study, 45 PD patients were followed for 12 months by motor UPDRS scores and FDG-PET imaging. The results showed a significant decrease in daily dyskinesias duration in STN AAV2-GAD group than that of the controls. Significant metabolic decreases in the striatum, thalamus, orbitofrontal, anterior cingulate, and prefrontal cortices were reported in the STN AAV2-GAD group compared with the sham group [61]. Although studies have emphasized the effectiveness of STN AAV2-GAD, it has been revealed that the mean level of improvement in the Deep Brain Stimulation (DBS) of the STN (STN DBS) was more than that observed with STN AAV2-GAD gene therapy [62].

Pe N rs ot on fo al rD U is se tri O bu n tio ly n

The recombinant AAA Vector (rAAV) is derived from the non-pathogenic parvoviruses facilitating proficient gene transfer into dividing and non-dividing cells in various organelles including the brains. Hadaczek and colleagues (2010) showed that AAV2-hAADC-mediated transgene expression continued for at least 8 years with no signs of neuroinflammation or reactive gliosis in a monkey’s brain [40]. Among AAVs 1–10, AAV-2-derived vectors are the most commonly used serotype in PD gene delivery, because a high percentage of gene expression is limited to the injected or manipulated regions such as SNc, and Subthalamic Nucleus (STN) [41, 42]. Dodiya and coworkers (2010) reported that AAV2/1 and AAV2/5 are superior to AAV2/8 for gene delivery to striatum of the nonhuman primate [43]. Moreover, it has been previously reported that the proficiency of AAV2/1, AAV2/5 is similar to or superior to that of AAV2/8 in the striatum [44-47]. While many vectors including the AAV have immune responses and cannot be reused if they fail, some of these disadvantages can overcome using hybrid vectors [48, 49].

Ghamgosha et al.

2.2. The Most Important Gene Therapy Approaches in PD

Currently, there are six most important gene therapy approaches for the treatment or improvement of the main motor signs of PD. Surprisingly, all of these gene therapy strategies have been developed based on AAV or lentivirus vectors. They consist of AAV-GAD, AAV-AADC, LentiAADC/TH/GTP-CH1, AAV-NRTN, α-Synuclein silencing, and PRKN gene delivery. AAV-GAD, AAV-AADC, and Lenti-AADC/TH/GTP-CH1 strategies improve motor signs via conveying neurotransmitter-producing enzyme and the remaining is based on gene delivery or disease-modifying approaches [50-53]. Development of the various gene therapy approaches is summarized in Fig. (2). 2.2.1. AAV-Glutamic Acid Decarboxylase (AAV-GAD)

PD is associated with a decrease in gamma-aminobutyric acid (GABA)ergic inputs into the STN and leads to its hyperactivity [54, 55]. Glutamic Acid Decarboxylase (GAD) is a central enzyme with two isoforms GAD65 and GAD67 and plays a key role in the production of GABA in the brain. GABA as the crucial inhibitory neurotransmitter of the Central Nervous System (CNS) compensates for the lost inhibitory mechanism in the basal ganglia via reducing neuronal excitability. Hence, transfer of the gene encoding GAD into the STN (STN AAV-GAD) or other basal ganglia nuclei could upsurge GABA synthesis and subsequently leads to an equilibrium of these pathways [50, 55, 56]. It has been shown that transfer of GAD-encoding gene into glutamatergic neurons of the STN using an AAV vector, leads to release of GABA in an activity-dependent manner in 6-hydroxydopamine (6-OHDA)-lesioned rats and consequently improve firing rates regulation and motor deficits

2.2.2. AAV–Aromatic L-Amino Acid Decarboxylase (AAV– AADC) Aromatic L-Amino Acid Decarboxylase (AADC) or DOPA decarboxylase is defined as a lyase enzyme that transforms pharmacologic or endogenous levodopa to dopamine, L-tryptophan to tryptamine, and 5-Hydroxytryptophan (5HTP) to serotonin. It has been suggested that the gradual death of dopaminergic neurons is associated with reduced

Therapeutic Advances in the Treatment of PD

PRKN gene delivery

a-Synuclein silencing

AAV-NRTN

Current Gene Therapy, 2018, Vol. 18, No. 4

209

Neuroprotective effect

Harmful potential

Insignificant improvement of the UPDRS score

Reduced dyskinesia

Lenti-AADC/TH/GTP-CH1

Minor improvements

AAV-AADC

Significant reduced dyskinesia

Pe N rs ot on fo al rD U is se tri O bu n tio ly n

AAV-GAD

Phase II

Phase I

Primate models

Rodent models

Fig. (2). Development of the various gene therapy approaches in PD. AAV-GAD: AAV-Glutamic Acid Decarboxylase, AAV-AADC: AAV–Aromatic L-Amino Acid Decarboxylase, Lenti-AADC/TH/GTP-CH1, Lenti-AADC/Tyrosine Hydroxylase/Guanosine Triphosphatecyclohydrolase I, AAV-NRTN: AAV–Neurturin. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this paper.)

activity of AADC [63]. Hence, transfer of the gene encoding AADC into STN (STN AAV-AADC) or other basal ganglia nuclei could increase AADC synthesis, consequently leading to enhancement of dopamine level [64]. Fan et al. stated that the simultaneous transduction of serotype-2 rAAV vector encoding AADC and Tyrosine Hydroxylase (TH) into the striatum (St rAAV2-AADC, and St rAAV2-TH) of 6-OHDA-lesioned rats enhanced L-dopa and dopamine synthesis more than St rAAV2-TH alone, and led to noteworthy decreases in apomorphine-induced rotational rate [65]. Parallel outcomes with an emphasis on increased dopamine production along with increased tetrahydrobiopterin (BH4) levels were achieved by the same team using a cocktail treatment of STN AAV-AADC, STN AAV-TH, and STN AAV-CH1 in 6-OHDA-lesioned rats [66]. Similarly, Azzouz and colleagues found that STN AAV-AADC, STN AAV-TH, as well as STN AAV-CH1 (STN AAV GTP cyclohydrolase 1) in a single lentiviral vector can lead to the constant expression of AADC, TH, CH1, and effective improvement of behavioral symptoms in 6-OHDA-lesioned rats; however, it is unfortunate that the expression of these enzymes in the striatum was detected for up to 5 months [23]. Furthermore, it has been revealed that delivery of the TH gene might be due to the overproduction of L-dopa thereby causing the delivery of higher degrees of dopamine, and eventually exacerbate motor disorders [67]. Unlike the suggestions of a different strategy using peripheral Ldihydroxyphenylalanine administration after STN AAVAADC by Leff and coworkers, detected transgene expression in the striatum for at least one year [68]. Some studies have shown the persistent expression of AADC with a 30% or more increases of the AADC tracer by [18F]fluoro-L-mtyrosine (FMT)-PET (FMT-PET) uptake in putamen or STN

AAV-AADC-treated individuals [22, 69-71], nonetheless, three cases of intracranial hemorrhages have also been reported [70]. Reduced levodopa dose, safety, and improvement of AADC activity with no neuroinflammation symptoms or reactive gliosis for at least 8 years are introduced as the advantages of STN AAV-AADC [40, 72]. It was suggested that the primary immunization with rAAV may limit the reduction of transduced gene expression in STN AAV2AADC-treated rats [73]. Unlike AAV2, other study stated that AAV-5 does not affect the humoral immune response [74]. Finally, the small size of the AAV (4kbp) is considered as another drawback which restricts the transmission of DNA to around 4 kb. 2.2.3. Lenti-AADC/TH/GTP-CH1 AADC, TH, and GTP-CH1 (Guanosine TriphosphateCyclohydrolase I) are three enzymes expressed in nigral neurons and play the crucial role in the synthesis of dopamine from tyrosine. AADC, TH, and GTP-CH1 catalyze L-DOPA to dopamine, amino acid L-tyrosine to L-3,4dihydroxyphenylalanine (L-DOPA), and Guanosine Triphosphate (GTP) to 7,8-dihydroneopterin triphosphate (DHNTP), respectively. Then, 6-pyruvoyltetrahydropterin synthase (PTPS), aldose reductase, and Sepiapterin Reductase (SR) catalyzes DHNTP, 6-Pyruvoyl-Tetrahydropterin (PTP), and sepiapterin to PTP, sepiapterin, and eventually BH4 as the cofactor of TH, respectively [75-77]. Therefore, transfer of the gene encoding AADC, TH, and GTP-CH1 into STN (STN AAV-AADC, STN AAV-TH, and STN AAV-GTPCH1) can induce ectopic dopamine synthesis and may be able to compensate for the amount of lost dopamine. As previously mentioned, Fan et al. (1998), Shen et al. (2000), and Azzouz and colleagues (2002) demonstrated the

210 Current Gene Therapy, 2018, Vol. 18, No. 4

treatment for PD without any clinical or systemic neurotoxicity [90, 91]. In a double-blind, phase 2 randomized trial, intraputamenal delivery of CERE-120 in 58 patients with advanced PD not only did not improve the motor UPDRS scores but also led to serious unwanted events including five developed tumors [21]. The same team also reported the beneficial effects of intraputamenal CERE-120 on the offmedication motor UPDRS scores and dyskinesia in 2008 [92]. Moreover, delivery of a single dose Nerve Growth Factor (NGF) (AAV2-NGF; CERE-110) and CERE-120 led to targeted-expression of bioactive protein in the monkeys and continued for at least 20 months for NRTN without any safety issues or antibody response to the protein [93]. Postmortem investigation has confirmed the lack of NRTN expression in the SNc in PD, whereas strong expression of NRTN and increased TH immunoreactivity in all monkeys following intraputamenal CERE-120(Put CERE120) treatment [94]. Alone or simultaneously, SNc or STN CERE-120 (SNc CERE-120 or STN CERE-120) delivery was conducted in 6-OHDA-lesioned rats and the results indicated that the delivery of SNc CERE-120 alone is sufficient in the regeneration and production of greater numbers of dopaminergic neurons and distribution of NRTN within the targeted nigrostriatal neurons [95]. On the other hand, it has been emphasized that bilateral Put or SNc plus Put CERE120 delivery is possible and safe [96, 97]. The effectiveness of bilateral treatment with SNc and Put CERE-120 in advanced patients with PD were compared with sham surgery group in a randomized, double-blind, and controlled trial. Data analysis showed that the SNc and Put CERE-120 delivery were not superior to sham surgery [98].

Pe N rs ot on fo al rD U is se tri O bu n tio ly n

beneficial effects of STN AAV-AADC, STN AAV-TH, and STN AAV-GTP-CH1 in 6-OHDA-lesioned rats [23, 65, 66]. Furthermore, it has been revealed that the Cos7 cells expressing AADC, TH, and GTP-CH1 and intrastriatal transduction of these genes (STN AADC/TH/GTP-CH1) enhance dopamine production in vivo and ex vivo [78, 79]. In 2009, Jarraya et al. illustrated increased dopamine production and noteworthy behavioral recovery in MPTP-lesioned macaque monkeys using intrastriatal delivery of a tricistronic lentiviral vector based on Equine Infectious Anemia Virus (EIAV) encoding AADC, TH, and GTP-CH1 in a single vector (EIAV-AADC-TH-GTP-CH1; named ProSavin), which continued up to 12 months without related dyskinesias [80]. In the following, an open-label, dose escalation, uncontrolled, phase 1/2 trial was conducted in 15 PD patients in third stages of the Hoehn & Yahr classification via ProSavin. All bilateral and intrastriatal ProSavin (STN ProSavin)-treated patients received three doses (low, mid, and high) in separate cohorts and were assessed with UPDRS (part III; off medication) scores. Significant improvement in motor UPDRS scores without any notable adverse events was observed in all patients at 6 and 12 months compared with baseline [81]. It was also claimed that ProSavin is safe and tolerable in patients with advanced PD.

Ghamgosha et al.

2.2.4. AAV–GDNF and AAV–NRTN (CERE-120)

Glial-derived Neurotrophic Factor (GDNF) and Neurturin (NRTN) are two neurotrophic factors with common functional abilities used to promote the survival and differentiation of nigrostriatal dopaminergic neurons both in vitro and in vivo [82-85]. It has been suggested that NRTN can promote the axonal regeneration, striatal dopamine levels, and reduce the methamphetamine-induced rotational rate in the 6-OHDA-lesioned rats [26]. Although GDNF and NRTN have common structural and functional similarities, direct delivery of each one into lateral cerebral ventricles leads to differential effects in vivo. GDNF-treated animals showed increased levels of dopamine utilization in the ventrolateral and mediodorsal striatum, whereas NRTN-treated animals showed increased levels only in the ventrolateral striatum. Weight loss and allodynia are also reported with GDNF administration but neither weight loss nor allodynia was detected with NRTN [86]. Kordower and colleagues (2006) hypothesized that the delivery of NRTN gene into the STN and SNc can preserve motor function in MPTP-induced hemiparkinsonian monkeys. They stated that the infusion of AAV2 vector encoding human NRTN (AAV–NRTN; also named CERE-120) can improve motor deficiencies by 80 to 90% (Since 4 to 10 months) and preserve nigrostriatal dopaminergic neurons [87]. Similar results were achieved by the CERE-120 delivery into the caudate and putamen in aged rhesus monkeys model of nigrostriatal dopamine deficiency [88]. CERE-120 delivery to the striatum (STN CERE-120) protected nigrostriatal dopaminergic neurons and improved motor function in 6-OHDA-lesioned rats for at least 6 months [89]. Parallel outcomes were realized by the same team using bilateral STN CERE-120 delivery. Herzog and coworkers (2007 and 2009) stated that STN CERE-120 delivery could enhance long-lasting expression (1-year) of NRTN and TH within the STN and SNc of rhesus monkeys. They also revealed that CERE-120 could be a novel safe

2.2.5. α-Synuclein Silencing

As mentioned previously, α-syn is known as the chief element of LBs encoded by the SNCA gene and is a chief player in both sporadic and familial cases of PD. The overexpression of α-syn leads to protein misfolding and its aggregation, dopaminergic cell loss and eventually LBs formation [99, 100]. Hence, downregulation or normalization of the α-syn level is considered to be a strategy for treatment of PD. On the contrary, delivery of α-syn into SNc leads to loss of nigral dopaminergic neurons, reduction of dopamine and TH levels, and ultimately the emergence of Parkinson-like neurodegeneration [101-103]. Similarly, it has been revealed that the duplication and triplication of α-syn locus cause PD [104, 105]. Interestingly, the dinucleotide repeat sequence (REP1) allele-length variation of the SNCA gene promoter is associated with susceptibility to PD [106]. Delivery of α-syn ribozyme via an AVV vector (AVV-αsynRz) into the SNc of MPTP-lesioned rats considerably protected TH-positive degeneration [107]. Similarly, silencing ectopic expression of SNCA using AAV-SNCA with or without a short hairpin RNA (shRNA) was conducted by infusion into rat SNc. Despite the behavioral improvement in the treated groups, high doses of shRNA-SNCA were toxic to dopaminergic neurons [108]. It has been shown that the RNA interference (RNAi) effectively diminish the expression of α-syn in both in vitro and in vivo [109-111]. Sapru et al. stated that the human and rat α-syn silencing is achievable in vitro and in vivo using lentiviral-mediated RNAi

Therapeutic Advances in the Treatment of PD

[112]. McCormack and colleagues designed a small interfering RNA (siRNA) against α-syn (siRNA-α-syn) and infused it into the left SNc of squirrel monkeys. From their results, significant (40-50%) suppression of α-syn was observed in the left vs. untreated right hemisphere, and siRNA-α-syn did not cause neuroinflammation or reduction of striatal dopamine concentrations [113]. Another study introduced a newly developed siRNA system which confers a moderate level of gene silencing by the name of expression-control RNAi (ExCont-RNAi). This team claimed that ExContRNAi-treated PD patients showed noteworthy improvement in their motor scores and announced ExCont-RNAi as a novel therapeutic method for PD with SNCA overexpression [114]. This is contrary to the report that RNAi-mediated αsyn silencing in the SNc leads to motor dysfunction and degeneration of nigral dopaminergic cells [115, 116]. 2.2.6. PRKN Gene Delivery

211

been shown that targeted correction of LRRK2 G2019S mutation in hiPSCs ameliorates neurite outgrowth, basal autophagy, and the increased α-syn and Microtubule-Associated Protein Tau (MAPT) expression in the mutated cells [142]. Likewise, Dansithong and colleagues (2015) showed that the missense mutation of SNCA gene can be genetically corrected by ZFNs in vitro [143]. It has also been demonstrated that LRRK2 mutant iPSC-derived dopaminergic neurons are more susceptible to oxidative stress than normal neurons [144, 145]. Liu et al. (2012) differentiated iPSC lines from patients with the Lamin A (LMNA) Mutation into Neural Stem Cells (NSCs) with or without homologous recombination-based gene correction. They observed that progressive nuclear aberrations of the PD-derived NSCs were absent in the corrected NSCs [146]. 3. STEM CELL-BASED THERAPIES FOR PD Over the past decades, stem cell-based therapies have made an increasingly smart therapeutic possibility to investigate and treat neurodegenerative diseases [147-150]. Replacing the lost cells is one of the main goals of this strategy. In this regard, replacing the dopamine-producing cells could help in the treatment or improvement of the disease. It was shown years ago that intra-parietal cortex transplantation of ventral mesencephalic tegmentum is able to generate new dopaminergic neurons in the 6-OHDA-lesioned rats [151]. However, the first clinical trial in candidates with PD was not performed by the human fetal mesencephalic tissue. Preclinical studies demonstrating that the human mesencephalic dopaminergic neurons from 6.5-8 week old fetuses engrafted to the non-immunosuppressed 6-OHDA-lesioned rats may be sufficient to renovate and reinnervate the striatum and improve behavioral deficits [152].

Pe N rs ot on fo al rD U is se tri O bu n tio ly n

PRKN is known as a key gene involved in autosomal recessive juvenile Parkinsonism, familial, and sporadic PD [117, 118]. PRKN also plays a crucial role in autophagy and the selective deletion of damaged mitochondria, consequently hindering the elimination of impaired mitochondria by the PD patients [119]. Paterna and coworkers observed that the SNc AAV-PRKN in MPTP-lesioned mice causes more living dopaminergic neurons and rotational rate improvement than the control [120]. Yasuda et al. also used the same strategy in MPTP-lesioned mice and stated that SNc AAV-PRKN delivery prevents motor deficits and dopaminergic cell loss [121]. Moreover, SNc Lenti-PRKN, two weeks prior to striatal 6-OHDA lesioning, led to the remarkable protection of dopaminergic cells and improved amphetamine-induced rotation rate [122]. Likewise, SNc LentiPRKN delivery caused overexpression of PRKN in an α-syn rat model of PD thereby significantly preserving TH-positive cell bodies and terminals in SNc and St, respectively [123]. Yamada and colleagues stated similar results with SNc rAAV2–PRKN and emphasized the effectiveness of this strategy against alpha-synucleinopathy [124]. Subsequently, the same team failed to repeat the protective effects of STN rAAV1–PRKN delivery in macaque monkeys, nonetheless, these conflicting findings may be related to serotypes (AAV1 vs. AAV2) or the injected brain regions (SNc vs. STN) [125].

Current Gene Therapy, 2018, Vol. 18, No. 4

2.2.7. Genome Editing Genome editing provides a powerful procedure to development of isogenic disease models and the repair of diseaserelated mutations, in vitro. Though studies have been reported the possibility of performing genome editing in human pluripotent stem cells (hPSCs) with Zinc Finger Nucleases (ZFNs) and Transcription Activator-like Effector Nucleases (TALENs) systems [126-135], Clustered Regularly Interspaced Short Palindromic Repeats (CRISPRs) system especially CRISPR/Cas9 protein (CRISPR/Cas9) is known as a simple and adaptive tool for modeling human cells for such purposes [136-141]. Soldner et al. (2011) generated isogenic stem cell models of PD by correction of an SNCAA53T mutation in PD patient-derived human induced pluripotent Stem Cells (hiPSCs) or by knock-in of this mutation in a Human Embryonic Stem Cell (hESC) line [131]. It has

3.1. Different Sources of Stem Cells for the Treatment of PD In recent years, several sources of stem cells have been established for the treatment of PD (Fig. 3), each of which has several advantages and disadvantages (Table 1). Embryonic Stem Cells (ESCs), Mesenchymal Stem Cells (MSCs), induced Pluripotent Stem Cells (iPSCs), fetal Ventral Mesencephalic Tissue (VMT; contains nigral dopaminergic neurons), and Neural Stem Cells (NSCs) are the most common sources that hold a great hope for the regeneration of the lost dopaminergic cells [153-156]. Each of these cell sources has some advantages and disadvantages that bring about specific results. Risk of tumor formation and genomic instability are the main limitations of using ESCs and iPSCs. In contrast, MSCs have slight immunogenic responses due to lack of Major Histocompatibility Complex class II (MHC-II) molecules [157, 158]. 3.1.1. Mesenchymal Stem Cells (MSCs) MSCs are a series of multipotent stem cells which are generally isolated from bone marrow [159-165], placenta [166], amniotic fluid [167], umbilical cord blood [168, 169], and adipose tissue [170-174]. MSCs can differentiate into adipocytes, neuron, skeletal myocytes, osteoblasts, tenocytes, chondrocytes, as well as the visceral mesoderm cells. These cells possess low tumorigenicity and immune

blastocyst

Ghamgosha et al.

Transplantation

212 Current Gene Therapy, 2018, Vol. 18, No. 4

hESCs

Transplantation

Biobsy

NSCs/Neurons

Somatic cell

Fetal brain

BM

NSCs/Neurons

MSC

iPSCs

Pe N rs ot on fo al rD U is se tri O bu n tio ly n

patient

Fig. (3). Different sources of stem cells for the treatment of PD. Table 1.

The potential for differentiation, advantages and disadvantages of different sources of stem cells for the treatment of PD.

Stem Cell

The Potential of Differentiation

Advantage

Disadvantage

NSCs

Neurons, Astrocytes, and Oligodendrocytes

1.

Low tumorigenicity and immune responses than ESCs and iPSCs

1.

Risk of graft-induced dyskinesias (GIDs)

2.

Limited differentiation in vivo

2.

Able to form neurons, astrocytes, and oligodendrocytes

3.

PD-like symptoms

4.

Failure to differentiate into the three germ layer cells

1.

Few therapeutic effectiveness in humans

2.

Failure to differentiate into the three germ layer cells

MSCs

ESCs

iPSCs

Adipocytes, osteoblasts, tenocytes, chondrocytes, skeletal myocytes, neuron, and the visceral mesoderm cells

Three germ layer cells

Three germ layer cells

1.

Low tumorigenicity and immune responses than ESCs and iPSCs

2.

A genuine and accessible cell source

3.

Improvement of cognitive and motor performances

1.

High proliferative and differential potency

1.

Risk of tumor formation

2.

Can generate dopaminergic neurons

2.

Genomic instability

3.

Able to differentiate into the three germ layer cells

3.

Immunogenic responses and ethical issues

1.

High proliferative and differential potency

1.

Risk of tumor formation

2.

Can generate dopaminergic neurons

2.

Genomic instability

3.

Able to differentiate into the three germ layer cells

4.

No immunogenic responses and ethical issues

responses than ESCs and iPSCs due to lack of MHC-II molecules [157, 158]. Although MSCs lack the limitations of ESCs and iPSCs, they have relatively few therapeutic effectiveness in humans. However, MSCs have been accepted as

promising candidates for cell therapy. Some reports have emphasized on the neuroprotective and neurorestorative beneficial effects of MSCs such as Bone Marrow-Derived MSCs (BMSCs) and adipose-derived mesenchymal stromal

Therapeutic Advances in the Treatment of PD

Current Gene Therapy, 2018, Vol. 18, No. 4

213

germ layer cells including dopaminergic neurons [147, 205, 206]. These unique features of ESCs make them an outstanding candidate for the generation of dopamine-producing cells. However, due to some limitations such as tumorigenicity and immunogenic responses, the use of undifferentiated forms of these cells has many problems. Even though low doses of undifferentiated mouse ESCs (mESCs) leads to developed severe teratoma-like tumors, some animals died before completion of behavioral evaluation [207]. A series of studies have reported the efficient induction of hESCs and mESCs into dopaminergic neurons [208-212]. Lee and coworkers (2000) presented an efficient method for the derivation of dopaminergic and serotonergic neurons from mESCs in vitro [208]. Kawasaki and colleagues (2000) showed that PA6 stromal cells promote neural differentiation probably by suppressing the effects of Bone Morphogenetic Protein 4 (BMP4) as an antineuralizing morphogen [209]. Other researchers have also tried to improve in vitro and/or in vivo induction or optimization of dopaminergic neurons from ESCs [213-219]. Likewise, it has been demonstrated that BMP inhibitor Noggin increases transformation rate of hESCs to dopaminergic neurons on PA6 stromal cells [220]. Although ESCs have the high proliferative potency to differentiate into the three germ layer cells including dopaminergic neurons which lead to significant improvement of clinical symptoms, genomic instability, immunogenic responses and the possibility of graft rejection, tumorigenicity and ethical issues are the limitations of ESCs. Hence, due to these drawbacks, no clinical trials have been accomplished yet with ESCs in PD patients.

3.1.2. Neural Stem Cells (NSCs)

3.1.4. Induced Pluripotent Stem Cells (iPSCs)

NSCs are multipotent, self-renewing stem cells that give rise to different neural cells such as neurons, dopaminergic neurons, oligodendrocytes, and astrocytes [189-192]. NSCs can be harvested from the fetal, neonatal, or adult brain regions such as SVZ and the Subgranular Zone (SGZ) of the dentate gyrus, as well as the Subependymal Zone (SEZ) of the lateral ventricles [193-196]. Clinical trials have shown that bilateral intraputamenal transplantation of VMT-derived dopaminergic neurons leads to some clinical benefits in younger but not in older PD patients, however, GIDs and off medication have also been reported as side effects [197, 198]. Moreover, long-term follow-up of two PD patients who had long-term (11-16 years) survival of transplanted VMT has been shown α-syn positive LBs in grafted neurons [199]. NSCs can be directly transplanted through the needle into a specific region like striatum or SNc. Nonetheless, BBB is an effective barrier to intravascular transplantation of NSCs [200]. Behavioral improvement has been shown in MPTP-lesioned Parkinsonian primates following injection of NSCs into the right SNc and the right and left caudate nuclei [201]. Intrastriatal transplantation of NSCs into the intact or 6-OHDA-lesioned rats led to the marked expression of Aromatic L-Amino Decarboxylase (AADC), while a significant number of the cells did not express TH [202]. Overall, NSCs exhibits low proliferation and differentiation capacities and lose their features after continuous passages [203, 204].

Since iPSCs were introduced by Nobel Prize-winning Shinya Yamanaka in August 2006, a tremendous transformation occurred in cell biology [221]. Afterward, Oct4, Sox2, Klf4, and c-Myc pluripotency factors were known as Yamanaka factors for induction of adult somatic cells to a novel type of pluripotent stem via retroviral or lentiviral systems cells called iPSCs [221-223]. Subsequently, Thomson and colleagues (2007) illustrated another protocol (OCT4, SOX2, NANOG, and LIN28) to induction of human somatic cells to iPSCs that exhibit the important characteristics of ESCs [224]. Genome-wide analysis with histone H3K27me3 and H3K4me3 modifications revealed that a slight difference exists between iPSCs and ESCs, and these cell lines are extremely similar to each other, however, the problem of the immunogenic response is almost resolved in iPSCs [225]. Thereafter, several studies used these two protocols to produce iPSCs and their differentiation into the dopaminergic neurons for the treatment or improvement of motor and nonmotor symptoms of PD [220, 226-234]. Some studies have also revealed that the autologous transplantation of iPSCsderived dopaminergic neurons can survive in the brain without immunosuppression and/or improve motor and nonmotor symptoms of PD [228, 231-236]. Although iPSCs have had almost slight problem in ethical issues, immunogenic responses and the possibility of graft rejection, several problems remain to be solved. First, it has been shown that the ectopic overexpression of Klf4, Oct4, Sox2, and c-Myc can lead to formation of breast tumors [237], dysplasia in epithelial cells [238], mucinous colon carcinoma [239], and

Pe N rs ot on fo al rD U is se tri O bu n tio ly n

cells [171-173], though contradictory results have also been reported [175]. It has been shown that grafted human MSCs (hMSCs) possess protective effects against nigrostriatal degeneration in human and 6-OHDA-lesioned rats [176-184]. Blandini and colleagues (2010) demonstrated that grafted hMSCs lead to increased number of striatal dopaminergic ⁺ terminals and TH nigral cell bodies in 6-OHDA-lesioned rats, thereby protecting nigrostriatal degeneration [176]. The neuroprotective effects and rescue of dopaminergic neurons were presented in MPTP-lesioned models [161, 185]. It has been shown that intravenous transplantation of MSCs can lead to inactivation of microglia and inhibition of dopaminergic cells death in MPTP-lesioned mice [186]. It has also been revealed that intrastriatal transplantation of dopaminergic-like cells derived from BMSCs successfully reduced the apomorphine-induced rotational rate in 6-OHDAlesioned mice. These cells had survived four months posttransplantation in the striatum and expressed TH, and some of them immigrated toward the SNc [187]. Furthermore, it is elucidated that hMSCs graft meaningfully augmented neurogenesis in both the Subventricular Zone (SVZ) and SNc of MPTP- and 6-OHDA-lesioned animal models, and lead to increased differentiation of neural precursor cells into dopaminergic neurons in the SNc [161, 188]. Conversely, Neirinckx and coworkers (2013) have stated that BMSCs are not able to replace lost neurons in MPTP-lesioned mice [175]. Despite considerable beneficial effects of MSCs transplantation in improving motor and non-motor symptoms in PD models, however, the underlying mechanisms have not been fully understood.

3.1.3. Embryonic Stem Cells (ESCs) ESCs are inner cell mass-derived pluripotent stem cells with high proliferative potency to differentiate into the three

214 Current Gene Therapy, 2018, Vol. 18, No. 4

the formation of human cancers [240], respectively. Second, 209 of 593 iPSCs genes were expressed in tumor and cancer tissues. Third, 5 oncogenes especially RAB25 were overexpressed in the iPSCs and iPSC-differentiated cells, respectively [241]. Therefore, tumorigenicity and formation of cancers are associated with genes. Furthermore, RAB25 encourages survival of cancer cells under various stress situations [242]. Another concern on this issue is the combination of viral structures within the host genome which can lead to teratoma formation [243]. 3.1.5. Directly Induced Dopaminergic Neurons (iDNs)

4. BRAIN STIMULATION 4.1. Transcranial Magnetic Stimulation (TMS) TMS is a noninvasive electromagnetic stimulation that has recently been used for the induction of neuroplasticity of the cerebral cortex for the improvement of neurological and psychiatric disorders like PD [261-279], Alzheimer’s disease (AD) [280-285], stroke [286-290], depression [291-294], schizophrenia (Scz) [295-298], obsessive-compulsive disorder (OCD) [299-301], epilepsy [302-306], and mood disorder [307-313]. Single pulse TMS releases a single pulse at a certain time, while repetitive pulse TMS (rTMS) deliver repeated single pulse to a specific brain region [314]. Some reports have stated the beneficial effects of rTMS on cognitive, psychological, or motor performances in PD patients [261, 272-274, 276, 277, 315-318], whereas, others have reported neutral (neither beneficial nor adverse effects) [264, 278] or adverse [269, 271] effects of rTMS. For the first time, Pascual-Leone and colleagues (1994) reported that subthreshold 5 Hz rTMS over the motor cortex improved the reaction time and movement time without affecting the rate of error in six patient with PD [261]. It has been elucidated that 1 and 5 Hz rTMS over the Motor hand area (M1) can improve dyskinesias and bradykinesia in PD patients, respectively [265, 274, 318]. Mally and coworkers (2004) concluded that 1 Hz rTMS for 7 days can postpone the progression of the disease than those treated with selegiline + levodopa [316]. The long-lasting motor UPDRS scores improvement has also been reported by rTMS in PD patients [266, 267, 274, 277, 317], and walking performance [272]. The motor improvement and the reduced methamphetamineinduced rotational rate have also been approved in 6-OHDAlesioned rats [276]. Conversely, it has been revealed that subthreshold 5-Hz rTMS over the motor cortex does not reliable or actually cause therapeutic effects on the movement of PD patients [262]. In addition, the cerebellar-like terminal and postural tremor may be induced by TMS in normal individuals [263]. Also, it has been demonstrated that low and/or high-frequency rTMS has no therapeutic effects and fail to improve motor performance in patients with PD [264, 278]. Abnormal M1 responses and subclinical worsening of the preliminary movement have been reported following rTMS over the primary motor cortex and supplementary motor area, respectively [269, 271]. A headache, ailment and burn at the site of stimulation, neck pain, seizure or convulsive syncope, mania, delusions, and vagal reactions are the reported side effects of rTMS [319-335]. However, further studies are required to identify the exact cellular and molecular mechanisms of rTMS.

Pe N rs ot on fo al rD U is se tri O bu n tio ly n

Transdifferentiation or lineage switching (as a class of metaplasia; cell switching from a specific fate to another) is defined as direct transformation of a somatic cell to another without passing a pluripotent phase [244, 245]. It was a great success so that fibroblasts would be directly reprogrammed to the target cells. It has been previously illustrated that mouse and human fibroblasts can be directly transformed to functional dopaminergic neurons without leaving a pluripotent phase [246-253]. A series of these studies have reported successful production of dopaminergic neurons from the mouse and/or human fibroblasts using different combinations of Mash1 (Ascl1), Myt1l, Nurr1 (Nr4a2), Brn2, Ngn2, Lmx1a, Sox2, and Pitx3 transcription factors [246, 250, 252, 254]. Park and coworkers (2006) used Mash1, Mytl1, Brn2, Lmx1a, Foxa2 transcription factors for the induction of fibroblasts to dopaminergic cells [254]. Caiazzo and colleagues (2011) showed that Mash1, Nurr1, and Lmx1a transcription factors can directly induce mouse and human fibroblasts to functional dopaminergic neurons [246]. Similar results have been reported by Pfisterer et al. (2011) using Ascl1, Brn2, and Myt1l transcription factors [252]. Liu et al. (2012) also identified that Mash1, Ngn2, Sox2, Nurr1, and Pitx3 transcription factors can efficiently convert human fibroblasts into dopaminergic-like cells [250]. In contrast, it has been revealed that Ascl1 and Pitx3 are sufficient in the transformation of fibroblasts into an immature dopaminergic cell fate [247]. Although this technique is faster compared to iPSC-derived dopaminergic neurons, several difficulties remain to be solved. It has been elucidated that sufficient numbers (200,000~400,000) of TH-positive cells are required to obtain a relatively acceptable result [255-260]. The iDNs produced by the procedure of Park et al. are not able to express all the specific markers of midbrain dopaminergic neurons and release the dopamine. Additionally, the efficacy of produced neurons has been reported to be about 10% [254]. The iDNs produced by Pfisterer group could secrete dopamine, however, their iDNs phenotype was immature and the efficacy of the produced neurons was also low [252]. Moreover, not all the specific markers of midbrain dopaminergic neurons could be expressed. The phenotype of iDNs produced by Caiazzo team was more similar to midbrain dopaminergic neurons. However, the transcriptome study of these iDNs, as in the two previous groups, showed a major difference with the midbrain-derived dopaminergic neurons [246]. As another limitation and safety concern, the fibroblasts of PD patients may carry the genomic mutations such as PRKN, SNCA, GBA, and LRRK2.

Ghamgosha et al.

4.2. Deep Brain Stimulation (DBS) DBS is an invasive neurosurgical technique introduced in 1987 as a clinical handling for the treatment of movement disorders, particularly for moderating the progressive symptoms of PD [1, 336-338]. DBS has now been extensively known as an effective method in advanced stages of PD. Usually, the Subthalamic Nucleus (STN), Ventral Intermediate nucleus of the thalamus (VIM), and Globus Pallidus Interna (GPi) are the key targets for DBS in PD [1]. It has previously been revealed that unilateral subthalamotomy including the caudal zona incerta can lead to significant improve-

Therapeutic Advances in the Treatment of PD

215

PRKN, stem cell-based therapies, M1 rTMS, STN DBS, GPi DBS, VIM DBS, and M1 tDCS, it seems that these hopes are not sufficient to treat the disease. As previously mentioned, the mean level of improvement by STN DBS was more than that seen with STN AAV2GAD gene therapy. Nevertheless, DBS as an invasive neurosurgical technique is done in severe stages of the disease, when more than 80% of dopaminergic neurons were lost. The small size of the AAV is considered as another drawback which restricts the transmission of DNA to around 4 kb. It has been described that RNAi-mediated α-syn silencing in the SNc leads to the motor dysfunction and degeneration of nigral dopaminergic cells. Although ESCs and iPSCs have high proliferative potency to differentiate into the three germ layer cells including dopaminergic neurons, genomic instability, immunogenic responses and the possibility of graft rejection, tumorigenicity and ethical issues are the limitations of these cell lines. Also, the fibroblasts of PD patients may carry genomic mutations such as PRKN, SNCA, GBA, and LRRK2. Although MSCs do not have the limitations of ESCs and iPSCs, they have relatively few therapeutic effectiveness in humans. Several breakdowns and side effects were described for rTMS and DBS. tDCS seems to be less complicated than rTMS and DBS, however, its effectiveness is also lower. Altogether, the effective treatment for PD has mostly failed because of these limitations and the late diagnosis of the disease. A day when scientists can acquire enough knowledge to prevent and treat this disease is sincerely hoped for.

Pe N rs ot on fo al rD U is se tri O bu n tio ly n

ment of contralateral motor symptoms in patients with advanced PD [339]. Subsequently, attempts of Blomstedt and colleagues (2009) led to important findings regarding the potential of the posterior subthalamic area including the caudal zona incerta as a DBS target in PD [340, 341]. It has been frequently shown that DBS of STN, GPi, and VIM (STN DBS, GPi DBS, and VIM DBS, respectively) can lead to relieve of resting and/or action tremors in patients with PD [340, 342-352]. Furthermore, constant improvements of bradykinesia, rigidity, and motor symptoms were also reported following STN and/or GPi DBS [353-358]. Although constant improvements in tremor, bradykinesia, rigidity, and motor symptoms have been stated following STN, VIM, or GPi DBS, however a series of complications have also been reported. Pneumocephalus, epileptic seizure, balance instability, anisocoria, hematoma, dyspnea, airway obstruction, infection, arterial hypertension and hypotension, bradycardia, tachycardia, gait disorders, cognitive decline, speech difficulty, and depression have been reported as the failures and side effects of DBS [351, 359-363].

Current Gene Therapy, 2018, Vol. 18, No. 4

4.3. Transcranial Direct Current Stimulation (tDCS)

Brain stimulation via feeble direct current has been shown to modulate excitability and plasticity of cortical and subcortical tissues [364, 365]. The use of feeble direct stimulation by applying galvanic currents date back to 1804 [366368]. tDCS is a non-invasive, safe, cheap, painless, and welltolerated brain stimulation technique that utilizes low direct current to stimulate a specific area of the brain [369]. While anodal tDCS acts to excite neural activity, cathodal tDCS inhibits or diminishes its activity. The therapeutic potential and long-lasting improvement of motor UPDRS scores, reaction time, bradykinesia, and dyskinesia have been reported by anodal tDCS of the primary motor cortex (M1) [370-377]. It has also been revealed that anodal tDCS of the DLPFC leads to improved executive functions including working memory in patients with PD [378, 379]. Tanaka et al. stated that cathodal but not anodal tDCS can significantly increase extracellular dopamine levels in the rat striatum [380]. It has been demonstrated that anodal but not cathodal tDCS of the motor cortex improves survival and integration of dopaminergic neurons in 6-OHDA-lesioned rats [381]. Although tDCS has some side effects like other brain stimulation techniques, these effects are relatively mild, transient, and less than the other methods. Some reports have presented a series of side effects like headache, nausea, and insomnia after tDCS [382, 383] however, not all, but some of the side effects may be due to the protocol used and the user's precision. CONCLUSION In this review, a detailed description of the cellular, molecular, and clinical therapeutic advances for the treatment of PD was presented and the advantages and disadvantages of these methods were emphasized. Despite the very respectable advances in these cutting-edge therapies, numerous problems still exist that has incapacitated the definitive treatment of the disease. Although numerous hopeful results emphasize the effectiveness of STN AAV2-GAD, STN AAV-AADC, STN AAV-TH, STN AAV-GTP-CH1, STN ProSavin, STN CERE-120, siRNA-α-syn, SNc rAAV2–

CONSENT FOR PUBLICATION Not applicable.

CONFLICT OF INTEREST The authors declare no conflict of interest, financial or otherwise. ACKNOWLEDGEMENTS

This review was supported by a grant from the Neuroscience Research Center, Baqiyatallah University of Medical Sciences, Tehran, Iran (grant no. BMSU.961010). REFERENCES [1] [2]

[3] [4] [5] [6] [7]

Mohammadi A, Mehdizadeh A. Deep brain stimulation and gene expression alterations in parkinson's disease. J Biomed Phys Eng 2016; 6(2): 47-50. Mohammadi A, Amooeian VG, Rashidi E. Dysfunction in brainderived neurotrophic factor signaling pathway and susceptibility to schizophrenia, parkinson’s and alzheimer’s diseases. Current Gene Therapy 2018; 18(1): 45-63. Javadpour A, Mohammadi A. Improving brain magnetic resonance image (MRI) segmentation via a novel algorithm based on genetic and regional growth. J Biomed Phys Eng 2016; 6(2): 95-108. Javadpour A, Mohammadi A. Implementing a smart method to eliminate artifacts of vital signals. J Biomed Phys Eng 2015; 5(4): 199-206. Klein C, Westenberger A. Genetics of parkinson's disease. Cold Spring Harb Perspect Med 2012; 2(1): a008888. Mizuta I, Satake W, Nakabayashi Y, et al. Multiple candidate gene analysis identifies alpha-synuclein as a susceptibility gene for sporadic Parkinson's disease. Hum Mol Genet 2006; 15(7): 1151-8. Nuytemans K, Theuns J, Cruts M, et al. Genetic etiology of Parkinson disease associated with mutations in the SNCA, PARK2,

216 Current Gene Therapy, 2018, Vol. 18, No. 4

[9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23]

[24] [25]

[26]

[27]

[28]

[29]

[30] [31]

[32]

[33]

[34] [35] [36]

Nutt JG, Burchiel KJ, Comella CL, et al. Randomized, doubleblind trial of glial cell line-derived neurotrophic factor (GDNF) in PD. Neurology 2003; 60(1): 69-73. Eriksdotter Jonhagen M, Nordberg A, Amberla K, et al. Intracerebroventricular infusion of nerve growth factor in three patients with Alzheimer's disease. Dement Geriatr Cogn Disord 1998; 9(5): 24657. Kordower JH, Palfi S, Chen EY, et al. Clinicopathological findings following intraventricular glial-derived neurotrophic factor treatment in a patient with Parkinson's disease. Ann Neurol 1999; 46(3): 419-24. Day-Lollini PA, Stewart GR, Taylor MJ, et al. Hyperplastic changes within the leptomeninges of the rat and monkey in response to chronic intracerebroventricular infusion of nerve growth factor. Exp Neurol 1997; 145(1): 24-37. Walther W, Stein U. Viral vectors for gene transfer: a review of their use in the treatment of human diseases. Drugs 2000; 60(2): 249-71. Thomas CE, Ehrhardt A, Kay MA. Progress and problems with the use of viral vectors for gene therapy. Nat Rev Genet 2003; 4(5): 346-58. Lundberg C, Bjorklund T, Carlsson T, et al. Applications of lentiviral vectors for biology and gene therapy of neurological disorders. Curr Gene Ther 2008; 8(6): 461-73. Jakobsson J, Lundberg C. Lentiviral vectors for use in the central nervous system. Mol Ther 2006; 13(3): 484-93. Matrai J, Chuah MK, VandenDriessche T. Recent advances in lentiviral vector development and applications. Mol Ther 2010; 18(3): 477-90. Ciesielska A, Hadaczek P, Mittermeyer G, et al. Cerebral infusion of AAV9 vector-encoding non-self proteins can elicit cell-mediated immune responses. Mol Ther 2013; 21(1): 158-66. Hadaczek P, Eberling JL, Pivirotto P, et al. Eight years of clinical improvement in MPTP-lesioned primates after gene therapy with AAV2-hAADC. Mol Ther 2010; 18(8): 1458-61. 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; 209(1): 58-71. Fitzsimons HL, Riban V, Bland RJ, et al. Biodistribution and safety assessment of AAV2-GAD following intrasubthalamic injection in the rat. J Gene Med 2010; 12(4): 385-98. Dodiya HB, Bjorklund T, Stansell J, et al. Differential transduction following basal ganglia administration of distinct pseudotyped AAV capsid serotypes in nonhuman primates. Mol Ther 2010; 18(3): 579-87. Burger C, Gorbatyuk OS, Velardo MJ, et al. Recombinant AAV viral vectors pseudotyped with viral capsids from serotypes 1, 2, and 5 display differential efficiency and cell tropism after delivery to different regions of the central nervous system. Mol Ther 2004; 10(2): 302-17. McFarland NR, Lee JS, Hyman BT, et al. Comparison of transduction efficiency of recombinant AAV serotypes 1, 2, 5, and 8 in the rat nigrostriatal system. J Neurochem 2009; 109(3): 838-45. Paterna JC, Feldon J, Bueler H. Transduction profiles of recombinant adeno-associated virus vectors derived from serotypes 2 and 5 in the nigrostriatal system of rats. J Virol 2004; 78(13): 6808-17. Taymans JM, Vandenberghe LH, Haute CV, et al. Comparative analysis of adeno-associated viral vector serotypes 1, 2, 5, 7, and 8 in mouse brain. Hum Gene Ther 2007; 18(3): 195-206. Nayak S, Herzog RW. Progress and prospects: immune responses to viral vectors. Gene Ther 2010; 17(3): 295-304. Zhou HS, Liu DP, Liang CC. Challenges and strategies: the immune responses in gene therapy. Med Res Rev 2004; 24(6): 74861. Emborg ME, Carbon M, Holden JE, et al. Subthalamic glutamic acid decarboxylase gene therapy: changes in motor function and cortical metabolism. J Cereb Blood Flow Metab 2007; 27(3): 5019. Luo J, Kaplitt MG, Fitzsimons HL, et al. Subthalamic GAD gene therapy in a Parkinson's disease rat model. Science 2002; 298(5592): 425-9. Aron L, Klein R. Repairing the parkinsonian brain with neurotrophic factors. Trends Neurosci 2011; 34(2): 88-100. Peterson AL, Nutt JG. Treatment of Parkinson's disease with trophic factors. Neurotherapeutics 2008; 5(2): 270-80.

Pe N rs ot on fo al rD U is se tri O bu n tio ly n

[8]

PINK1, PARK7, and LRRK2 genes: a mutation update. Hum Mutat 2010; 31(7): 763-80. Camargos ST, Dornas LO, Momeni P, et al. Familial Parkinsonism and early onset Parkinson's disease in a Brazilian movement disorders clinic: phenotypic characterization and frequency of SNCA, PRKN, PINK1, and LRRK2 mutations. Mov Disord 2009; 24(5): 662-6. Sidransky E, Nalls MA, Aasly JO, et al. Multicenter analysis of glucocerebrosidase mutations in Parkinson's disease. N Engl J Med 2009; 361(17): 1651-61. Malgieri G, Eliezer D. Structural effects of Parkinson's disease linked DJ-1 mutations. Protein Sci 2008; 17(5): 855-68. Polymeropoulos MH, Higgins JJ, Golbe LI, et al. Mapping of a gene for Parkinson's disease to chromosome 4q21-q23. Science 1996; 274(5290): 1197-9. Dodson MW, Guo M. Pink1, Parkin, DJ-1 and mitochondrial dysfunction in Parkinson's disease. Curr Opin Neurobiol 2007; 17(3): 331-7. Thomas KJ, McCoy MK, Blackinton J, et al. DJ-1 acts in parallel to the PINK1/parkin pathway to control mitochondrial function and autophagy. Hum Mol Genet 2011; 20(1): 40-50. Shi M, Furay A, Sossi V, et al. DJ-1 and αSYN in LRRK2 CSF do not correlate with striatal dopaminergic function. Neurobiol Aging 2012; 33(4): 836.e5-836.e7. Cookson MR. Parkinsonism Due to Mutations in PINK1, parkin, and dj-1 and oxidative stress and mitochondrial pathways. Cold Spring Harbor Perspect Med 2012; 2(9): a009415. Waragai M, Sekiyama K, Sekigawa A, et al. alpha-Synuclein and DJ-1 as potential biological fluid biomarkers for Parkinson's Disease. Int J Mol Sci 2010; 11(11): 4257-66. Cookson MR. α-Synuclein and neuronal cell death. Mole Neurodegenerat 2009; 4(1): 9. Kim T, Vemuganti R. Mechanisms of Parkinson’s disease-related proteins in mediating secondary brain damage after cerebral ischemia. J Cerebral Blood Flow Metabol 2017; 37(6): 1910-26. Adler CH. Nonmotor complications in Parkinson's disease. Move Disord 2005; 20(S11): S23-S29. Kordower JH, Emborg ME, Bloch J, et al. Neurodegeneration prevented by lentiviral vector delivery of GDNF in primate models of Parkinson's disease. Science 2000; 290(5492): 767-73. Marks WJ, Jr., Bartus RT, Siffert J, et al. Gene delivery of AAV2neurturin for Parkinson's disease: a double-blind, randomised, controlled trial. Lancet Neurol 2010; 9(12): 1164-1172. Eberling JL, Jagust WJ, Christine CW, et al. Results from a phase I safety trial of hAADC gene therapy for Parkinson disease. Neurology 2008; 70(21): 1980-3. Azzouz M, Martin-Rendon E, Barber RD, et al. Multicistronic lentiviral vector-mediated striatal gene transfer of aromatic Lamino 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; 22(23): 10302-12. Rosenblad C, Kirik D, Bjorklund A. Neurturin enhances the survival of intrastriatal fetal dopaminergic transplants. Neuroreport 1999; 10(8): 1783-7. Kozlowski DA, Connor B, Tillerson JL, et al. Delivery of a GDNF gene into the substantia nigra after a progressive 6-OHDA lesion maintains functional nigrostriatal connections. Exp Neurol 2000; 166(1): 1-15. Oiwa Y, Yoshimura R, Nakai K, et al. Dopaminergic neuroprotection and regeneration by neurturin assessed by using behavioral, biochemical and histochemical measurements in a model of progressive Parkinson's disease. Brain Res 2002; 947(2): 271-83. Kirik D, Rosenblad C, Bjorklund A. Preservation of a functional nigrostriatal dopamine pathway by GDNF in the intrastriatal 6OHDA lesion model depends on the site of administration of the trophic factor. Eur J Neurosci 2000; 12(11): 3871-82. Grondin R, Cass WA, Zhang Z, et al. Glial cell line-derived neurotrophic factor increases stimulus-evoked dopamine release and motor speed in aged rhesus monkeys. J Neurosci 2003; 23(5): 1974-80. Wang L, Muramatsu S, Lu Y, et al. Delayed delivery of AAVGDNF prevents nigral neurodegeneration and promotes functional recovery in a rat model of Parkinson's disease. Gene Ther 2002; 9(6): 381-9.

Ghamgosha et al.

[37] [38] [39] [40] [41] [42] [43]

[44]

[45] [46] [47] [48] [49] [50]

[51] [52] [53]

Therapeutic Advances in the Treatment of PD

[55] [56] [57] [58]

[59] [60] [61] [62] [63] [64]

[65]

[66]

[67] [68]

[69] [70] [71] [72] [73] [74] [75] [76] [77]

Bergman H, Wichmann T, DeLong MR. Reversal of experimental parkinsonism by lesions of the subthalamic nucleus. Science 1990; 249(4975): 1436-8. Hamani C, Saint-Cyr JA, Fraser J, et al. The subthalamic nucleus in the context of movement disorders. Brain 2004; 127(Pt 1): 4-20. LeWitt PA, Rezai AR, Leehey MA, et al. AAV2-GAD gene therapy for advanced Parkinson's disease: a double-blind, sham-surgery controlled, randomised trial. Lancet Neurol 2011; 10(4): 309-19. Lee B, Lee H, Nam YR, et al. Enhanced expression of glutamate decarboxylase 65 improves symptoms of rat parkinsonian models. Gene Ther 2005; 12(15): 1215-22. Kaplitt MG, Feigin A, Tang C, et al. Safety and tolerability of gene therapy with an adeno-associated virus (AAV) borne GAD gene for Parkinson's disease: an open label, phase I trial. Lancet 2007; 369(9579): 2097-105. Feigin A, Kaplitt MG, Tang C, et al. Modulation of metabolic brain networks after subthalamic gene therapy for Parkinson's disease. Proc Natl Acad Sci U S A 2007; 104(49): 19559-64. Feigin A, Eidelberg D. Gene transfer therapy for neurodegenerative disorders. Mov Disord 2007; 22(9): 1223-8. Niethammer M, Tang CC, LeWitt PA, et al. Long-term follow-up of a randomized AAV2-GAD gene therapy trial for Parkinson's disease. JCI Insight 2017; 2(7): e90133. Deuschl G, Schade-Brittinger C, Krack P, et al. A randomized trial of deep-brain stimulation for Parkinson's disease. N Engl J Med 2006; 355(9): 896-908. Nagatsu T, Sawada M. Biochemistry of postmortem brains in Parkinson's disease: historical overview and future prospects. J Neural Transm Suppl 2007(72): 113-20. Bankiewicz KS, Eberling JL, Kohutnicka M, et al. Convectionenhanced 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. Fan DS, Ogawa M, Fujimoto KI, et al. Behavioral recovery in 6hydroxydopamine-lesioned rats by cotransduction of striatum with tyrosine hydroxylase and aromatic L-amino acid decarboxylase genes using two separate adeno-associated virus vectors. Hum Gene Ther 1998; 9(17): 2527-35. Shen Y, Muramatsu SI, Ikeguchi K, et al. Triple transduction with adeno-associated virus vectors expressing tyrosine hydroxylase, aromatic-L-amino-acid decarboxylase, and GTP cyclohydrolase I for gene therapy of Parkinson's disease. Hum Gene Ther 2000; 11(11): 1509-19. Carlsson T, Bjorklund T, Kirik D. Restoration of the striatal dopamine synthesis for Parkinson's disease: viral vector-mediated enzyme replacement strategy. Curr Gene Ther 2007; 7(2): 109-20. Leff SE, Spratt SK, Snyder RO, et al. Long-term restoration of striatal L-aromatic amino acid decarboxylase activity using recombinant adeno-associated viral vector gene transfer in a rodent model of Parkinson's disease. Neuroscience 1999; 92(1): 185-96. Muramatsu S, Fujimoto K, Kato S, et al. A phase I study of aromatic L-amino acid decarboxylase gene therapy for Parkinson's disease. Mol Ther 2010; 18(9): 1731-5. Christine CW, Starr PA, Larson PS, et al. Safety and tolerability of putaminal AADC gene therapy for Parkinson disease. Neurology 2009; 73(20): 1662-9. Forsayeth JR, Eberling JL, Sanftner LM, et al. A dose-ranging study of AAV-hAADC therapy in Parkinsonian monkeys. Mol Ther 2006; 14(4): 571-7. Bankiewicz KS, Forsayeth J, Eberling JL, et al. Long-term clinical improvement in MPTP-lesioned primates after gene therapy with AAV-hAADC. Mol Ther 2006; 14(4): 564-70. Sanftner LM, Suzuki BM, Doroudchi MM, et al. Striatal delivery of rAAV-hAADC to rats with preexisting immunity to AAV. Mol Ther 2004; 9(3): 403-9. Abeliovich A, Schmitz Y, Farinas I, et al. Mice lacking alphasynuclein display functional deficits in the nigrostriatal dopamine system. Neuron 2000; 25. Kumer SC, Vrana KE. Intricate regulation of tyrosine hydroxylase activity and gene expression. J Neurochem 1996; 67(2): 443-62. Nagatsu T, Ichinose H. GTP cyclohydrolase I gene, dystonia, juvenile parkinsonism, and Parkinson's disease. J Neural Transm Suppl 1997; 49: 203-9. Elsworth JD, Roth RH. Dopamine synthesis, uptake, metabolism, and receptors: relevance to gene therapy of Parkinson's disease. Exp Neurol 1997; 144(1): 4-9.

[78]

[79]

[80] [81]

[82] [83]

[84] [85] [86] [87]

[88]

[89]

[90] [91]

[92]

[93]

[94]

[95]

[96] [97] [98]

217

Duan CL, Su Y, Zhao CL, et al. The assays of activities and function of TH, AADC, and GCH1 and their potential use in ex vivo gene therapy of PD. Brain Res Brain Res Protoc 2005; 16(1-3): 3743. Lu LL, Su Y, Duan CL, et al. [Gene therapy of tyrosine hydroxylase, aromatic L-amino acid decarboxylase, and GTP cyclohydrolase genes in rat model of Parkinson's disease]. Zhonghua Yi Xue Za Zhi 2004; 84(18): 1528-32. Jarraya B, Boulet S, Ralph GS, et al. Dopamine gene therapy for Parkinson's disease in a nonhuman primate without associated dyskinesia. Sci Transl Med 2009; 1(2): 2ra4. Palfi S, Gurruchaga JM, Ralph GS, et al. Long-term safety and tolerability of ProSavin, a lentiviral vector-based gene therapy for Parkinson's disease: a dose escalation, open-label, phase 1/2 trial. Lancet 2014; 383(9923): 1138-46. Kotzbauer PT, Lampe PA, Heuckeroth RO, et al. Neurturin, a relative of glial-cell-line-derived neurotrophic factor. Nature 1996; 384(6608): 467-70. Horger BA, Nishimura MC, Armanini MP, et al. Neurturin Exerts Potent Actions on Survival and Function of Midbrain Dopaminergic Neurons. The Journal of Neuroscience 1998; 18(13): 49294937. Lin LF, Doherty DH, Lile JD, et al. GDNF: a glial cell line-derived neurotrophic factor for midbrain dopaminergic neurons. Science 1993; 260(5111): 1130-2. Li H, He Z, Su T, et al. Protective action of recombinant neurturin on dopaminergic neurons in substantia nigra in a rhesus monkey model of Parkinson's disease. Neurol Res 2003; 25(3): 263-7. Hoane MR, Gulwadi AG, Morrison S, et al. Differential in vivo effects of neurturin and glial cell-line-derived neurotrophic factor. Exp Neurol 1999; 160(1): 235-43. Kordower JH, Herzog CD, Dass B, et al. Delivery of neurturin by AAV2 (CERE-120)-mediated gene transfer provides structural and functional neuroprotection and neurorestoration in MPTP-treated monkeys. Ann Neurol 2006; 60(6): 706-15. Herzog CD, Dass B, Holden JE, et al. Striatal delivery of CERE120, an AAV2 vector encoding human neurturin, enhances activity of the dopaminergic nigrostriatal system in aged monkeys. Mov Disord 2007; 22(8): 1124-32. Gasmi M, Brandon EP, Herzog CD, 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; 27(1): 67-76. Herzog CD, Dass B, Gasmi M, et al. Transgene expression, bioactivity, and safety of CERE-120 (AAV2-neurturin) following delivery to the monkey striatum. Mol Ther 2008; 16(10): 1737-44. Herzog CD, Brown L, Gammon D, et al. Expression, bioactivity, and safety 1 year after adeno-associated viral vector type 2mediated delivery of neurturin to the monkey nigrostriatal system support cere-120 for Parkinson's disease. Neurosurgery 2009; 64(4): 602-12. Marks WJ, Jr., Ostrem JL, Verhagen L, 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; 7(5): 400-8. Herzog CD, Bishop KM, Brown L, et al. Gene transfer provides a practical means for safe, long-term, targeted delivery of biologically active neurotrophic factor proteins for neurodegenerative diseases. Drug Deliv Transl Res 2011; 1(5): 361-82. Bartus RT, Herzog CD, Chu Y, et al. Bioactivity of AAV2neurturin gene therapy (CERE-120): differences between Parkinson's disease and nonhuman primate brains. Mov Disord 2011; 26(1): 27-36. Herzog CD, Brown L, Kruegel BR, et al. Enhanced neurotrophic distribution, cell signaling and neuroprotection following substantia nigral versus striatal delivery of AAV2-NRTN (CERE-120). Neurobiol Dis 2013; 58: 38-48. Bartus RT, Baumann TL, Siffert J, et al. Safety/feasibility of targeting the substantia nigra with AAV2-neurturin in Parkinson patients. Neurology 2013; 80(18): 1698-701. Marks WJ. Jr; Baumann TL, Bartus RT. Long-Term Safety of Patients with Parkinson's Disease Receiving rAAV2-Neurturin (CERE-120) Gene Transfer. Hum Gene Ther 2016; 27(7): 522-7. Warren Olanow C, Bartus RT, Baumann TL, et al. Gene delivery of neurturin to putamen and substantia nigra in Parkinson disease:

Pe N rs ot on fo al rD U is se tri O bu n tio ly n

[54]

Current Gene Therapy, 2018, Vol. 18, No. 4

218 Current Gene Therapy, 2018, Vol. 18, No. 4

[100] [101] [102]

[103] [104] [105] [106] [107]

[108]

[109] [110] [111] [112] [113] [114]

[115] [116] [117] [118] [119] [120] [121]

[122] [123]

[124] [125] [126]

[127]

[128]

Vercammen L, Van der Perren A, Vaudano E, et al. Parkin protects against neurotoxicity in the 6-hydroxydopamine rat model for Parkinson's disease. Mol Ther 2006; 14(5): 716-23. Lo Bianco C, Schneider BL, Bauer M, et al. Lentiviral vector delivery of parkin prevents dopaminergic degeneration in an alphasynuclein rat model of Parkinson's disease. Proc Natl Acad Sci U S A 2004; 101(50): 17510-5. Yamada M, Mizuno Y, Mochizuki H. Parkin gene therapy for alpha-synucleinopathy: a rat model of Parkinson's disease. Hum Gene Ther 2005; 16(2): 262-70. Yasuda T, Miyachi S, Kitagawa R, et al. Neuronal specificity of alpha-synuclein toxicity and effect of Parkin co-expression in primates. Neuroscience 2007; 144. Lombardo A, Genovese P, Beausejour CM, et al. Gene editing in human stem cells using zinc finger nucleases and integrasedefective lentiviral vector delivery. Nat Biotechnol 2007; 25(11): 1298-306. Suzuki K, Mitsui K, Aizawa E, et al. Highly efficient transient gene expression and gene targeting in primate embryonic stem cells with helper-dependent adenoviral vectors. Proc Natl Acad Sci U S A 2008; 105(37): 13781-6. Zou J, Mali P, Huang X, et al. Site-specific gene correction of a point mutation in human iPS cells derived from an adult patient with sickle cell disease. Blood 2011; 118(17): 4599-608. Hockemeyer D, Soldner F, Beard C, et al. Efficient targeting of expressed and silent genes in human ESCs and iPSCs using zincfinger nucleases. Nat Biotechnol 2009; 27(9): 851-7. Hockemeyer D, Wang H, Kiani S, et al. Genetic engineering of human pluripotent cells using TALE nucleases. Nat Biotechnol 2011; 29(8): 731-4. Soldner F, Laganiere J, Cheng AW, et al. Generation of isogenic pluripotent stem cells differing exclusively at two early onset Parkinson point mutations. Cell 2011; 146(2): 318-31. Yusa K, Rashid ST, Strick-Marchand H, et al. Targeted gene correction of alpha1-antitrypsin deficiency in induced pluripotent stem cells. Nature 2011; 478(7369): 391-4. Zou J, Maeder ML, Mali P, et al. Gene targeting of a diseaserelated gene in human induced pluripotent stem and embryonic stem cells. Cell Stem Cell 2009; 5(1): 97-110. Sebastiano V, Maeder ML, Angstman JF, et al. In situ genetic correction of the sickle cell anemia mutation in human induced pluripotent stem cells using engineered zinc finger nucleases. Stem Cells 2011; 29(11): 1717-26. Li M, Suzuki K, Qu J, et al. Efficient correction of hemoglobinopathy-causing mutations by homologous recombination in integration-free patient iPSCs. Cell Res 2011; 21(12): 1740-4. Horvath P, Barrangou R. CRISPR/Cas, the immune system of bacteria and archaea. Science 2010; 327(5962): 167-70. Wang H, Yang H, Shivalila CS, et al. One-step generation of mice carrying mutations in multiple genes by CRISPR/Cas-mediated genome engineering. Cell 2013; 153(4): 910-8. Mali P, Yang L, Esvelt KM, et al. RNA-guided human genome engineering via Cas9. Science 2013; 339(6121): 823-6. Chen YC, Farzadfard F, Gharaei N, et al. Randomized crispr-cas transcriptional perturbation screening reveals protective genes against alpha-synuclein toxicity. Mol Cell 2017; 68(1): 247-257.e5. Soldner F, Stelzer Y, Shivalila CS, et al. Parkinson-associated risk variant in distal enhancer of alpha-synuclein modulates target gene expression. Nature 2016; 533(7601): 95-9. Arias-Fuenzalida J, Jarazo J, Qing X, et al. FACS-assisted crisprcas9 genome editing facilitates parkinson's disease modeling. Stem Cell Reports 2017; 9(5): 1423-1431. Reinhardt P, Schmid B, Burbulla LF, et al. Genetic correction of a LRRK2 mutation in human iPSCs links parkinsonian neurodegeneration to ERK-dependent changes in gene expression. Cell Stem Cell 2013; 12(3): 354-67. Dansithong W, Paul S, Scoles DR, et al. Generation of snca cell models using zinc finger nuclease (zfn) technology for efficient high-throughput drug screening. PLoS One 2015; 10(8): e0136930. Nguyen HN, Byers B, Cord B, et al. LRRK2 mutant iPSC-derived DA neurons demonstrate increased susceptibility to oxidative stress. Cell Stem Cell 2011; 8(3): 267-80. Heo HY, Park JM, Kim CH, et al. LRRK2 enhances oxidative stress-induced neurotoxicity via its kinase activity. Exp Cell Res 2010; 316(4): 649-56.

Pe N rs ot on fo al rD U is se tri O bu n tio ly n

[99]

A double-blind, randomized, controlled trial. Ann Neurol 2015; 78(2): 248-57. Bellucci A, Navarria L, Falarti E, et al. Redistribution of DAT/alpha-synuclein complexes visualized by in situ proximity ligation assay in transgenic mice modelling early Parkinson's disease. PLoS One 2011; 6(12): e27959. Kim C, Lee SJ. Controlling the mass action of alpha-synuclein in Parkinson's disease. J Neurochem 2008; 107(2): 303-16. Kirik D, Rosenblad C, Burger C, et al. Parkinson-like neurodegeneration induced by targeted overexpression of alpha-synuclein in the nigrostriatal system. J Neurosci 2002; 22(7): 2780-91. Yamada M, Iwatsubo T, Mizuno Y, et al. Overexpression of alphasynuclein in rat substantia nigra results in loss of dopaminergic neurons, phosphorylation of alpha-synuclein and activation of caspase-9: resemblance to pathogenetic changes in Parkinson's disease. J Neurochem 2004; 91(2): 451-61. Decressac M, Kadkhodaei B, Mattsson B, et al. alpha-Synucleininduced down-regulation of Nurr1 disrupts GDNF signaling in nigral dopamine neurons. Sci Transl Med 2012; 4(163): 163ra156. Singleton AB, Farrer M, Johnson J, et al. alpha-Synuclein locus triplication causes Parkinson's disease. Science 2003; 302(5646): 841. Nishioka K, Hayashi S, Farrer MJ, et al. Clinical heterogeneity of alpha-synuclein gene duplication in Parkinson's disease. Ann Neurol 2006; 59(2): 298-309. Maraganore DM, de Andrade M, Elbaz A, et al. Collaborative analysis of alpha-synuclein gene promoter variability and Parkinson disease. JAMA 2006; 296(6): 661-70. Hayashita-Kinoh H, Yamada M, Yokota T, et al. Down-regulation of alpha-synuclein expression can rescue dopaminergic cells from cell death in the substantia nigra of Parkinson's disease rat model. Biochem Biophys Res Commun 2006; 341(4): 1088-95. Khodr CE, Sapru MK, Pedapati J, et al. An alpha-synuclein AAV gene silencing vector ameliorates a behavioral deficit in a rat model of Parkinson's disease, but displays toxicity in dopamine neurons. Brain Res 2011; 1395: 94-107. Junn E, Lee KW, Jeong BS, et al. Repression of alpha-synuclein expression and toxicity by microRNA-7. Proc Natl Acad Sci U S A 2009; 106(31): 13052-7. Mak SK, McCormack AL, Manning-Bog AB, et al. Lysosomal degradation of alpha-synuclein in vivo. J Biol Chem 2010; 285(18): 13621-9. Lewis J, Melrose H, Bumcrot D, et al. In vivo silencing of alphasynuclein using naked siRNA. Mol Neurodegener 2008; 3: 19. Sapru MK, Yates JW, Hogan S, et al. Silencing of human alphasynuclein in vitro and in rat brain using lentiviral-mediated RNAi. Exp Neurol 2006; 198(2): 382-92. McCormack AL, Mak SK, Henderson JM, et al. Alpha-synuclein suppression by targeted small interfering RNA in the primate substantia nigra. PLoS One 2010; 5(8): e12122. Takahashi M, Suzuki M, Fukuoka M, et al. Normalization of overexpressed alpha-synuclein causing parkinson's disease by a moderate gene silencing with rna interference. Mol Ther Nucleic Acids 2015; 4: e241. Gorbatyuk OS, Li S, Nash K, et al. In vivo RNAi-mediated alphasynuclein silencing induces nigrostriatal degeneration. Mol Ther 2010; 18(8): 1450-7. Gorbatyuk OS, Li S, Nguyen FN, et al. Alpha-Synuclein expression in rat substantia nigra suppresses phospholipase D2 toxicity and nigral neurodegeneration. Mol Ther 2010; 18(10): 1758-68. Kitada T, Asakawa S, Hattori N, et al. Mutations in the parkin gene cause autosomal recessive juvenile parkinsonism. Nature 1998; 392(6676): 605-8. Lucking CB, Durr A, Bonifati V, et al. Association between earlyonset Parkinson's disease and mutations in the parkin gene. N Engl J Med 2000; 342(21): 1560-7. Narendra D, Tanaka A, Suen DF, et al. Parkin is recruited selectively to impaired mitochondria and promotes their autophagy. J Cell Biol 2008; 183(5): 795-803. Paterna JC, Leng A, Weber E, et al. DJ-1 and Parkin modulate dopamine-dependent behavior and inhibit MPTP-induced nigral dopamine neuron loss in mice. Mol Ther 2007; 15(4): 698-704. Yasuda T, Hayakawa H, Nihira T, et al. Parkin-mediated protection of dopaminergic neurons in a chronic MPTP-minipump mouse model of Parkinson disease. J Neuropathol Exp Neurol 2011; 70(8): 686-97.

Ghamgosha et al.

[129] [130] [131] [132] [133] [134]

[135] [136] [137] [138] [139] [140] [141] [142]

[143] [144] [145]

Therapeutic Advances in the Treatment of PD

[147] [148]

[149]

[150]

[151] [152]

[153] [154] [155] [156] [157] [158]

[159] [160]

[161]

[162]

[163] [164] [165] [166]

[167]

Liu GH, Suzuki K, Qu J, et al. Targeted gene correction of laminopathy-associated LMNA mutations in patient-specific iPSCs. Cell Stem Cell 2011; 8(6): 688-94. Mohammadi A, Attari F, Babapour V, et al. Generation of rat embryonic germ cells via inhibition of TGFβ and MEK pathways. Cell J 2015; 17(2): 288-295. Hosseini SR, Kaka G, Joghataei MT, et al. Assessment of neuroprotective properties of melissa officinalis in combination with human umbilical cord blood stem cells after spinal cord injury. ASN Neuro 2016; 8(6). Kaka G, Arum J, Sadraie SH, et al. Bone marrow stromal cells associated with poly l-lactic-co-glycolic acid (PLGA) nanofiber scaffold improve transected sciatic nerve regeneration. Iranian J Biotechnol 2017; 15(3): 149-156. Mohammadi A, Maleki-Jamshid A, Sanooghi D, et al. Transplantation of human chorion-derived cholinergic progenitor cells: a novel treatment for neurological disorders. Mol Neurobiol 2018: doi: 10.1007/s12035-018-0968-1. [Epub ahead of print]. Bjorklund A, Stenevi U. Reconstruction of the nigrostriatal dopamine pathway by intracerebral nigral transplants. Brain Res 1979; 177(3): 555-60. Brundin P, Strecker RE, Widner H, et al. Human fetal dopamine neurons grafted in a rat model of Parkinson's disease: immunological aspects, spontaneous and drug-induced behaviour, and dopamine release. Exp Brain Res 1988; 70(1): 192-208. Nam H, Lee KH, Nam DH, et al. Adult human neural stem cell therapeutics: Current developmental status and prospect. World J Stem Cells 2015; 7(1): 126-36. Herberts CA, Kwa MS, Hermsen HP. Risk factors in the development of stem cell therapy. J Transl Med 2011; 9: 29. Kitada M, Dezawa M. Parkinson's disease and mesenchymal stem cells: potential for cell-based therapy. Parkinsons Dis 2012; 2012: 873706. Politis M, Lindvall O. Clinical application of stem cell therapy in Parkinson's disease. BMC Med 2012; 10: 1. Eliopoulos N, Stagg J, Lejeune L, et al. Allogeneic marrow stromal cells are immune rejected by MHC class I- and class II-mismatched recipient mice. Blood 2005; 106(13): 4057-65. Romieu-Mourez R, Francois M, Boivin MN, et al. Regulation of MHC class II expression and antigen processing in murine and human mesenchymal stromal cells by IFN-gamma, TGF-beta, and cell density. J Immunol 2007; 179(3): 1549-58. Pittenger MF, Mackay AM, Beck SC, et al. Multilineage potential of adult human mesenchymal stem cells. Science 1999; 284(5411): 143-7. Venkataramana NK, Pal R, Rao SA, et al. Bilateral transplantation of allogenic adult human bone marrow-derived mesenchymal stem cells into the subventricular zone of Parkinson's disease: a pilot clinical study. Stem Cells Int 2012; 2012: 931902. Park HJ, Shin JY, Lee BR, et al. Mesenchymal stem cells augment neurogenesis in the subventricular zone and enhance differentiation of neural precursor cells into dopaminergic neurons in the substantia nigra of a parkinsonian model. Cell Transplant 2012; 21(8): 1629-40. Wang F, Yasuhara T, Shingo T, et al. Intravenous administration of mesenchymal stem cells exerts therapeutic effects on parkinsonian model of rats: focusing on neuroprotective effects of stromal cellderived factor-1alpha. BMC Neurosci 2010; 11: 52. Chen D, Fu W, Zhuang W, et al. Therapeutic effects of intranigral transplantation of mesenchymal stem cells in rat models of Parkinson's disease. J Neurosci Res 2017; 95(3): 907-917. Huang Y, Chang C, Zhang J, et al. Bone marrow-derived mesenchymal stem cells increase dopamine synthesis in the injured striatum. Neural Regen Res 2012; 7(34): 2653-62. Khoo MLM, Tao H, Meedeniya ACB, et al. Transplantation of Neuronal-Primed Human Bone Marrow Mesenchymal Stem Cells in Hemiparkinsonian Rodents. PLOS One 2011; 6(5): e19025. Zhang Y, Li C, Jiang X, et al. Human placenta-derived mesenchymal progenitor cells support culture expansion of long-term culture-initiating cells from cord blood CD34+ cells. Exp Hematol 2004; 32(7): 657-64. Roubelakis MG, Pappa KI, Bitsika V, et al. Molecular and proteomic characterization of human mesenchymal stem cells derived from amniotic fluid: comparison to bone marrow mesenchymal stem cells. Stem Cells Dev 2007; 16(6): 931-52.

[168] [169] [170] [171]

[172]

[173]

[174] [175]

[176]

[177] [178]

[179] [180]

[181] [182] [183] [184]

[185]

[186]

[187]

[188]

219

Erices A, Conget P, Minguell JJ. Mesenchymal progenitor cells in human umbilical cord blood. Br J Haematol 2000; 109(1): 235-42. Bieback K, Kern S, Kluter H, et al. Critical parameters for the isolation of mesenchymal stem cells from umbilical cord blood. Stem Cells 2004; 22(4): 625-34. Izadpanah R, Trygg C, Patel B, et al. Biologic properties of mesenchymal stem cells derived from bone marrow and adipose tissue. J Cell Biochem 2006; 99(5): 1285-97. Schwerk A, Altschuler J, Roch M, et al. Human adipose-derived mesenchymal stromal cells increase endogenous neurogenesis in the rat subventricular zone acutely after 6-hydroxydopamine lesioning. Cytotherapy 2015; 17(2): 199-214. Schwerk A, Altschuler J, Roch M, et al. Adipose-derived human mesenchymal stem cells induce long-term neurogenic and antiinflammatory effects and improve cognitive but not motor performance in a rat model of Parkinson's disease. Regen Med 2015; 10(4): 431-46. Berg J, Roch M, Altschuler J, et al. Human adipose-derived mesenchymal stem cells improve motor functions and are neuroprotective in the 6-hydroxydopamine-rat model for Parkinson's disease when cultured in monolayer cultures but suppress hippocampal neurogenesis and hippocampal memory function when cultured in spheroids. Stem Cell Rev 2015; 11(1): 133-49. Zuk PA, Zhu M, Mizuno H, et al. Multilineage cells from human adipose tissue: implications for cell-based therapies. Tissue Eng 2001; 7(2): 211-28. Neirinckx V, Marquet A, Coste C, et al. Adult Bone Marrow Neural Crest Stem Cells and Mesenchymal Stem Cells Are Not Able to Replace Lost Neurons in Acute MPTP-Lesioned Mice. PLOS One 2013; 8(5): e64723. Blandini F, Cova L, Armentero MT, et al. Transplantation of undifferentiated human mesenchymal stem cells protects against 6hydroxydopamine neurotoxicity in the rat. Cell Transplant 2010; 19(2): 203-17. Woodbury D, Schwarz EJ, Prockop DJ, et al. Adult rat and human bone marrow stromal cells differentiate into neurons. J Neurosci Res 2000; 61(4): 364-70. Li Y, Chen J, Wang L, et al. Intracerebral transplantation of bone marrow stromal cells in a 1-methyl-4-phenyl-1,2,3,6tetrahydropyridine mouse model of Parkinson's disease. Neurosci Lett 2001; 316(2): 67-70. Dezawa M, Kanno H, Hoshino M, et al. Specific induction of neuronal cells from bone marrow stromal cells and application for autologous transplantation. J Clin Invest 2004; 113(12): 1701-10. Honma T, Honmou O, Iihoshi S, et al. Intravenous infusion of immortalized human mesenchymal stem cells protects against injury in a cerebral ischemia model in adult rat. Exp Neurol 2006; 199(1): 56-66. Koh SH, Kim KS, Choi MR, et al. Implantation of human umbilical cord-derived mesenchymal stem cells as a neuroprotective therapy for ischemic stroke in rats. Brain Res 2008; 1229: 233-48. Park HJ, Lee PH, Bang OY, et al. Mesenchymal stem cells therapy exerts neuroprotection in a progressive animal model of Parkinson's disease. J Neurochem 2008; 107(1): 141-51. Lanza C, Morando S, Voci A, et al. Neuroprotective mesenchymal stem cells are endowed with a potent antioxidant effect in vivo. J Neurochem 2009; 110(5): 1674-84. Karussis D, Karageorgiou C, Vaknin-Dembinsky A, et al. Safety and immunological effects of mesenchymal stem cell transplantation in patients with multiple sclerosis and amyotrophic lateral sclerosis. Arch Neurol 2010; 67(10): 1187-94. Park HJ, Bang G, Lee BR, et al. Neuroprotective effect of human mesenchymal stem cells in an animal model of double toxininduced multiple system atrophy parkinsonism. Cell Transplant 2011; 20(6): 827-35. Chao YX, He BP, Tay SS. Mesenchymal stem cell transplantation attenuates blood brain barrier damage and neuroinflammation and protects dopaminergic neurons against MPTP toxicity in the substantia nigra in a model of Parkinson's disease. J Neuroimmunol 2009; 216(1-2): 39-50. Offen D, Barhum Y, Levy YS, et al. Intrastriatal transplantation of mouse bone marrow-derived stem cells improves motor behavior in a mouse model of Parkinson's disease. J Neural Transm Suppl 2007(72): 133-43. Cova L, Armentero MT, Zennaro E, et al. Multiple neurogenic and neurorescue effects of human mesenchymal stem cell after trans-

Pe N rs ot on fo al rD U is se tri O bu n tio ly n

[146]

Current Gene Therapy, 2018, Vol. 18, No. 4

220 Current Gene Therapy, 2018, Vol. 18, No. 4

[190] [191] [192] [193] [194] [195] [196] [197] [198] [199] [200] [201]

[202]

[203] [204] [205] [206] [207]

[208] [209] [210] [211] [212]

[213]

[214]

[215] [216] [217]

[218] [219] [220]

Morizane A, Takahashi J, Takagi Y, et al. Optimal conditions for in vivo induction of dopaminergic neurons from embryonic stem cells through stromal cell-derived inducing activity. J Neurosci Res 2002; 69(6): 934-9. Morizane A, Takahashi J, Shinoyama M, et al. Generation of graftable dopaminergic neuron progenitors from mouse ES cells by a combination of coculture and neurosphere methods. J Neurosci Res 2006; 83(6): 1015-27. Zeng X, Cai J, Chen J, et al. Dopaminergic differentiation of human embryonic stem cells. Stem Cells 2004; 22(6): 925-40. Sasai Y. Generation of dopaminergic neurons from embryonic stem cells. J Neurol 2002; 249 Suppl 2: Ii41-4. Shintani A, Nakao N, Kakishita K, et al. Generation of dopamine neurons from embryonic stem cells in the presence of the neuralizing activity of bone marrow stromal cells derived from adult mice. J Neurosci Res 2008; 86(13): 2829-38. Park CH, Minn YK, Lee JY, et al. In vitro and in vivo analyses of human embryonic stem cell-derived dopamine neurons. J Neurochem 2005; 92(5): 1265-76. Yamazoe H, Iwata H. Efficient generation of dopaminergic neurons from mouse embryonic stem cells enclosed in hollow fibers. Biomaterials 2006; 27(28): 4871-80. Chiba S, Lee YM, Zhou W, et al. Noggin enhances dopamine neuron production from human embryonic stem cells and improves behavioral outcome after transplantation into Parkinsonian rats. Stem Cells 2008; 26(11): 2810-20. Takahashi K, Yamanaka S. Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell 2006; 126(4): 663-76. Takahashi K, Tanabe K, Ohnuki M, et al. Induction of pluripotent stem cells from adult human fibroblasts by defined factors. Cell 2007; 131(5): 861-72. Takahashi K, Okita K, Nakagawa M, et al. Induction of pluripotent stem cells from fibroblast cultures. Nat Protoc 2007; 2(12): 3081-9. Yu J, Vodyanik MA, Smuga-Otto K, et al. Induced pluripotent stem cell lines derived from human somatic cells. Science 2007; 318(5858): 1917-20. Guenther MG, Frampton GM, Soldner F, et al. Chromatin structure and gene expression programs of human embryonic and induced pluripotent stem cells. Cell Stem Cell 2010; 7(2): 249-57. Cooper O, Hargus G, Deleidi M, et al. Differentiation of human ES and Parkinson's disease iPS cells into ventral midbrain dopaminergic neurons requires a high activity form of SHH, FGF8a and specific regionalization by retinoic acid. Mol Cell Neurosci 2010; 45(3): 258-66. Sanchez-Danes A, Richaud-Patin Y, Carballo-Carbajal I, et al. Disease-specific phenotypes in dopamine neurons from human iPSbased models of genetic and sporadic Parkinson's disease. EMBO Mol Med 2012; 4(5): 380-95. Sundberg M, Bogetofte H, Lawson T, et al. Improved cell therapy protocols for Parkinson's disease based on differentiation efficiency and safety of hESC-, hiPSC-, and non-human primate iPSC-derived dopaminergic neurons. Stem Cells 2013; 31(8): 1548-62. Sanchez-Danes A, Consiglio A, Richaud Y, et al. Efficient generation of A9 midbrain dopaminergic neurons by lentiviral delivery of LMX1A in human embryonic stem cells and induced pluripotent stem cells. Hum Gene Ther 2012; 23(1): 56-69. Doi D, Samata B, Katsukawa M, et al. Isolation of human induced pluripotent stem cell-derived dopaminergic progenitors by cell sorting for successful transplantation. Stem Cell Rep 2014; 2(3): 33750. Morizane A, Doi D, Kikuchi T, et al. Direct comparison of autologous and allogeneic transplantation of iPSC-derived neural cells in the brain of a non-human primate. Stem Cell Rep 2013; 1(4): 28392. Kikuchi T, Morizane A, Doi D, et al. Survival of human induced pluripotent stem cell-derived midbrain dopaminergic neurons in the brain of a primate model of Parkinson's disease. J Parkinsons Dis 2011; 1(4): 395-412. Samata B, Doi D, Nishimura K, et al. Purification of functional human ES and iPSC-derived midbrain dopaminergic progenitors using LRTM1. Nat Commun 2016; 7: 13097. Kikuchi T, Morizane A, Doi D, et al. Human iPS cell-derived dopaminergic neurons function in a primate Parkinson's disease model. Nature 2017; 548(7669): 592-6.

Pe N rs ot on fo al rD U is se tri O bu n tio ly n

[189]

plantation in an experimental model of Parkinson's disease. Brain Res 2010; 1311: 12-27. Temple S. The development of neural stem cells. Nature 2001; 414(6859): 112-7. Uchida N, Buck DW, He D, et al. Direct isolation of human central nervous system stem cells. Proc Natl Acad Sci U S A 2000; 97(26): 14720-5. Gage FH. Mammalian neural stem cells. Science 2000; 287(5457): 1433-8. Wagner J, Akerud P, Castro DS, et al. Induction of a midbrain dopaminergic phenotype in Nurr1-overexpressing neural stem cells by type 1 astrocytes. Nat Biotechnol 1999; 17(7): 653-9. Taupin P, Gage FH. Adult neurogenesis and neural stem cells of the central nervous system in mammals. J Neurosci Res 2002; 69(6): 745-9. Deierborg T, Soulet D, Roybon L, et al. Emerging restorative treatments for Parkinson's disease. Prog Neurobiol 2008; 85(4): 407-32. Kempermann G, Kuhn HG, Gage FH. Genetic influence on neurogenesis in the dentate gyrus of adult mice. Proceedings of the National Academy of Sciences 1997; 94(19): 10409-14. Luo J, Daniels SB, Lennington JB, et al. The aging neurogenic subventricular zone. Aging Cell 2006; 5(2): 139-52. Olanow CW, Goetz CG, Kordower JH, et al. A double-blind controlled trial of bilateral fetal nigral transplantation in Parkinson's disease. Ann Neurol 2003; 54(3): 403-14. Freed CR, Greene PE, Breeze RE, et al. Transplantation of embryonic dopamine neurons for severe Parkinson's disease. N Engl J Med 2001; 344(10): 710-9. Li JY, Englund E, Holton JL, et al. Lewy bodies in grafted neurons in subjects with Parkinson's disease suggest host-to-graft disease propagation. Nat Med 2008; 14(5): 501-3. Pendharkar AV, Chua JY, Andres RH, et al. Biodistribution of neural stem cells after intravascular therapy for hypoxic-ischemia. Stroke 2010; 41(9): 2064-70. Redmond DE, Jr., Bjugstad KB, Teng YD, et al. Behavioral improvement in a primate Parkinson's model is associated with multiple homeostatic effects of human neural stem cells. Proc Natl Acad Sci U S A 2007; 104(29): 12175-80. Yang M, Stull ND, Berk MA, et al. Neural stem cells spontaneously express dopaminergic traits after transplantation into the intact or 6-hydroxydopamine-lesioned rat. Exp Neurol 2002; 177(1): 50-60. Ramos-Moreno T, Lendinez JG, Pino-Barrio MJ, et al. Clonal human fetal ventral mesencephalic dopaminergic neuron precursors for cell therapy research. PLoS One 2012; 7(12): e52714. Villa A, Liste I, Courtois ET, et al. Generation and properties of a new human ventral mesencephalic neural stem cell line. Exp Cell Res 2009; 315(11): 1860-74. Thomson JA, Itskovitz-Eldor J, Shapiro SS, et al. Embryonic stem cell lines derived from human blastocysts. Science 1998; 282(5391): 1145-7. Evans MJ, Kaufman MH. Establishment in culture of pluripotential cells from mouse embryos. Nature 1981; 292(5819): 154-6. Bjorklund LM, Sanchez-Pernaute R, Chung S, et al. Embryonic stem cells develop into functional dopaminergic neurons after transplantation in a Parkinson rat model. Proc Natl Acad Sci U S A 2002; 99(4): 2344-9. Lee SH, Lumelsky N, Studer L, et al. Efficient generation of midbrain and hindbrain neurons from mouse embryonic stem cells. Nat Biotechnol 2000; 18(6): 675-9. Kawasaki H, Mizuseki K, Nishikawa S, et al. Induction of midbrain dopaminergic neurons from ES cells by stromal cell-derived inducing activity. Neuron 2000; 28(1): 31-40. Perrier AL, Tabar V, Barberi T, et al. Derivation of midbrain dopamine neurons from human embryonic stem cells. Proc Natl Acad Sci U S A 2004; 101(34): 12543-8. Kriks S, Shim JW, Piao J, et al. Dopamine neurons derived from human ES cells efficiently engraft in animal models of Parkinson's disease. Nature 2011; 480(7378): 547-51. Kim JH, Auerbach JM, Rodriguez-Gomez JA, et al. Dopamine neurons derived from embryonic stem cells function in an animal model of Parkinson's disease. Nature 2002; 418(6893): 50-6.

Ghamgosha et al.

[221] [222] [223] [224] [225] [226]

[227]

[228]

[229]

[230]

[231]

[232]

[233] [234]

Therapeutic Advances in the Treatment of PD

[236]

[237] [238] [239]

[240] [241]

[242]

[243]

[244] [245] [246] [247] [248] [249] [250] [251]

[252] [253] [254] [255]

[256] [257]

Emborg ME, Liu Y, Xi J, et al. Induced pluripotent stem cellderived neural cells survive and mature in the nonhuman primate brain. Cell Rep 2013; 3(3): 646-50. Redmond DE, Jr., Vinuela A, Kordower JH, et al. Influence of cell preparation and target location on the behavioral recovery after striatal transplantation of fetal dopaminergic neurons in a primate model of Parkinson's disease. Neurobiol Dis 2008; 29(1): 103-16. Ghaleb AM, Nandan MO, Chanchevalap S, et al. Kruppel-like factors 4 and 5: the yin and yang regulators of cellular proliferation. Cell Res 2005; 15(2): 92-6. Hochedlinger K, Yamada Y, Beard C, et al. Ectopic expression of Oct-4 blocks progenitor-cell differentiation and causes dysplasia in epithelial tissues. Cell 2005; 121(3): 465-77. Park ET, Gum JR, Kakar S, et al. Aberrant expression of SOX2 upregulates MUC5AC gastric foveolar mucin in mucinous cancers of the colorectum and related lesions. Int J Cancer 2008; 122(6): 1253-60. Kuttler F, Mai S. c-Myc, Genomic Instability and Disease. Genome Dyn 2006; 1: 171-90. Zhang G, Shang B, Yang P, et al. Induced pluripotent stem cell consensus genes: implication for the risk of tumorigenesis and cancers in induced pluripotent stem cell therapy. Stem Cells Dev 2012; 21(6): 955-64. Cheng K, Agarwal R, Mitra S, et al. Rab25 Small GTPase Mediates Secretion of Tumor Necrosis Factor Receptor Superfamily Member 11b (osteoprotegerin) Protecting Cancer Cells from Effects of TRAIL. J Genet Syndr Gene Ther 2013; 4. Howe SJ, Mansour MR, Schwarzwaelder K, et al. Insertional mutagenesis combined with acquired somatic mutations causes leukemogenesis following gene therapy of SCID-X1 patients. J Clin Invest 2008; 118(9): 3143-50. Davis RL, Weintraub H, Lassar AB. Expression of a single transfected cDNA converts fibroblasts to myoblasts. Cell 1987; 51(6): 987-1000. Slack JM. Metaplasia and transdifferentiation: from pure biology to the clinic. Nat Rev Mol Cell Biol 2007; 8(5): 369-78. Caiazzo M, Dell'Anno MT, Dvoretskova E, et al. Direct generation of functional dopaminergic neurons from mouse and human fibroblasts. Nature 2011; 476(7359): 224-7. Kim J, Su SC, Wang H, et al. Functional integration of dopaminergic neurons directly converted from mouse fibroblasts. Cell Stem Cell 2011; 9(5): 413-9. Dell'Anno MT, Caiazzo M, Leo D, et al. Remote control of induced dopaminergic neurons in parkinsonian rats. J Clin Invest 2014; 124(7): 3215-29. Kim HS, Kim J, Jo Y, et al. Direct lineage reprogramming of mouse fibroblasts to functional midbrain dopaminergic neuronal progenitors. Stem Cell Res 2014; 12(1): 60-8. Liu X, Li F, Stubblefield EA, et al. Direct reprogramming of human fibroblasts into dopaminergic neuron-like cells. Cell Res 2012; 22(2): 321-32. Liu X, Huang Q, Li F, et al. Enhancing the efficiency of direct reprogramming of human primary fibroblasts into dopaminergic neuron-like cells through p53 suppression. Sci China Life Sci 2014; 57(9): 867-75. Pfisterer U, Kirkeby A, Torper O, et al. Direct conversion of human fibroblasts to dopaminergic neurons. Proc Natl Acad Sci 2011; 108(25): 10343-8. Jiang H, Xu Z, Zhong P, et al. Cell cycle and p53 gate the direct conversion of human fibroblasts to dopaminergic neurons. Nat Commun 2015; 6: 10100. Park CH, Kang JS, Shin YH, et al. Acquisition of in vitro and in vivo functionality of Nurr1-induced dopamine neurons. Faseb j 2006; 20(14): 2553-5. Mendez I, Sanchez-Pernaute R, Cooper O, et al. Cell type analysis of functional fetal dopamine cell suspension transplants in the striatum and substantia nigra of patients with Parkinson's disease. Brain 2005; 128(Pt 7): 1498-510. Li JY, Englund E, Holton JL, et al. Lewy bodies in grafted neurons in subjects with Parkinson's disease suggest host-to-graft disease propagation. Nat Med 2008; 14(5): 501-3. Kordower JH, Rosenstein JM, Collier TJ, et al. Functional fetal nigral grafts in a patient with Parkinson's disease: chemoanatomic, ultrastructural, and metabolic studies. J Comp Neurol 1996; 370(2): 203-30.

[258] [259] [260] [261] [262] [263] [264]

[265] [266] [267]

[268] [269] [270] [271] [272]

[273]

[274]

[275]

[276] [277] [278]

[279] [280]

221

Hagell P, Schrag A, Piccini P, et al. Sequential bilateral transplantation in Parkinson's disease: effects of the second graft. Brain 1999; 122 ( Pt 6): 1121-32. Mendez I, Vinuela A, Astradsson A, et al. Dopamine neurons implanted into people with Parkinson's disease survive without pathology for 14 years. Nat Med 2008; 14(5): 507-9. Yang N, Ng YH, Pang ZP, et al. Induced neuronal cells: how to make and define a neuron. Cell Stem Cell 2011; 9(6): 517-25. Pascual-Leone A, Valls-Sole J, Brasil-Neto JP, et al. Akinesia in Parkinson's disease. II. Effects of subthreshold repetitive transcranial motor cortex stimulation. Neurology 1994; 44(5): 892-8. Ghabra MB, Hallett M, Wassermann EM. Simultaneous repetitive transcranial magnetic stimulation does not speed fine movement in PD. Neurology 1999; 52(4): 768-70. Topka H, Mescheriakov S, Boose A, et al. A cerebellar-like terminal and postural tremor induced in normal man by transcranial magnetic stimulation. Brain 1999; 122 ( Pt 8): 1551-62. Tergau F, Wassermann EM, Paulus W, et al. Lack of clinical improvement in patients with Parkinson's disease after low and high frequency repetitive transcranial magnetic stimulation. Electroencephalogr Clin Neurophysiol Suppl 1999; 51: 281-8. Siebner HR, Mentschel C, Auer C, et al. Repetitive transcranial magnetic stimulation has a beneficial effect on bradykinesia in Parkinson's disease. Neuroreport 1999; 10(3): 589-94. Mally J, Stone TW. Improvement in Parkinsonian symptoms after repetitive transcranial magnetic stimulation. J Neurol Sci 1999; 162(2): 179-84. Mally J, Stone TW. Therapeutic and dose-dependent effect of repetitive microelectroshock induced by transcranial magnetic stimulation in Parkinson's disease. J Neurosci Res 1999; 57(6): 935-40. Shimamoto H, Morimitsu H, Sugita S, et al. Therapeutic effect of repetitive transcranial magnetic stimulation in Parkinson's disease. Rinsho Shinkeigaku 1999; 39(12): 1264-7. Boylan LS, Pullman SL, Lisanby SH, et al. Repetitive transcranial magnetic stimulation to SMA worsens complex movements in Parkinson's disease. Clin Neurophysiol 2001; 112(2): 259-64. Filipovic SR, Bhatia KP, Rothwell JC. 1-Hz repetitive transcranial magnetic stimulation and diphasic dyskinesia in Parkinson's disease. Mov Disord 2013; 28(2): 245-6. Hanajima R, Terao Y, Shirota Y, et al. Triad-conditioning transcranial magnetic stimulation in Parkinson's disease. Brain Stimul 2014; 7(1): 74-9. Mak MK. Repetitive transcranial magnetic stimulation combined with treadmill training can modulate corticomotor inhibition and improve walking performance in people with Parkinson's disease. J Physiother 2013; 59(2): 128. Nardone R, De Blasi P, Holler Y, et al. Repetitive transcranial magnetic stimulation transiently reduces punding in Parkinson's disease: a preliminary study. J Neural Transm (Vienna) 2014; 121(3): 267-74. Sayin S, Cakmur R, Yener GG, et al. Low-frequency repetitive transcranial magnetic stimulation for dyskinesia and motor performance in Parkinson's disease. J Clin Neurosci 2014; 21(8): 1373-6. Zanjani A, Zakzanis KK, Daskalakis ZJ, et al. Repetitive transcranial magnetic stimulation of the primary motor cortex in the treatment of motor signs in Parkinson's disease: A quantitative review of the literature. Mov Disord 2015; 30(6): 750-8. Lee JY, Kim SH, Ko AR, et al. Therapeutic effects of repetitive transcranial magnetic stimulation in an animal model of Parkinson's disease. Brain Res 2013; 1537: 290-302. Maruo T, Hosomi K, Shimokawa T, et al. High-frequency repetitive transcranial magnetic stimulation over the primary foot motor area in Parkinson's disease. Brain Stimul 2013; 6(6): 884-91. Benninger DH, Iseki K, Kranick S, et al. Controlled study of 50-Hz repetitive transcranial magnetic stimulation for the treatment of Parkinson disease. Neurorehabil Neural Repair 2012; 26(9): 1096105. Edwards MJ, Talelli P, Rothwell JC. Clinical applications of transcranial magnetic stimulation in patients with movement disorders. Lancet Neurol 2008; 7(9): 827-40. Haffen E, Chopard G, Pretalli JB, et al. A case report of daily left prefrontal repetitive transcranial magnetic stimulation (rTMS) as an adjunctive treatment for Alzheimer disease. Brain Stimul 2012; 5(3): 264-6.

Pe N rs ot on fo al rD U is se tri O bu n tio ly n

[235]

Current Gene Therapy, 2018, Vol. 18, No. 4

222 Current Gene Therapy, 2018, Vol. 18, No. 4

[282] [283]

[284]

[285]

[286]

[287] [288]

[289] [290]

[291] [292]

[293] [294] [295] [296]

[297]

[298]

[299]

[300]

[301]

Ahmed MA, Darwish ES, Khedr EM, et al. Effects of low versus high frequencies of repetitive transcranial magnetic stimulation on cognitive function and cortical excitability in Alzheimer's dementia. J Neurol 2012; 259(1): 83-92. Ferreri F, Pasqualetti P, Maatta S, et al. Motor cortex excitability in Alzheimer's disease: a transcranial magnetic stimulation follow-up study. Neurosci Lett 2011; 492(2): 94-8. Eliasova I, Anderkova L, Marecek R, et al. Non-invasive brain stimulation of the right inferior frontal gyrus may improve attention in early Alzheimer's disease: a pilot study. J Neurol Sci 2014; 346(1-2): 318-22. Devi G, Voss HU, Levine D, et al. Open-label, short-term, repetitive transcranial magnetic stimulation in patients with Alzheimer's disease with functional imaging correlates and literature review. Am J Alzheimers Dis Other Demen 2014; 29(3): 248-55. Trebbastoni A, Gilio F, D'Antonio F, et al. Chronic treatment with rivastigmine in patients with Alzheimer's disease: a study on primary motor cortex excitability tested by 5 Hz-repetitive transcranial magnetic stimulation. Clin Neurophysiol 2012; 123(5): 902-9. Barros GSC, Borba Costa dSR, Borba dSP, et al. Efficacy of coupling repetitive transcranial magnetic stimulation and physical therapy to reduce upper-limb spasticity in patients with stroke: a randomized controlled trial. Arch Phys Med Rehabil 2014; 95(2): 222-9. Mang CS, Borich MR, Brodie SM, et al. Diffusion imaging and transcranial magnetic stimulation assessment of transcallosal pathways in chronic stroke. Clin Neurophysiol 2015; 126(10): 1959-71. Sasaki N, Mizutani S, Kakuda W, et al. Comparison of the effects of high- and low-frequency repetitive transcranial magnetic stimulation on upper limb hemiparesis in the early phase of stroke. J Stroke Cerebrovasc Dis 2013; 22(4): 413-8. Balyts'kyi OP. Using transcranial magnetic stimulation for studying functional state of motor centers in patients with ischemic stroke. Lik Sprava 2013(5): 75-80. Massie CL, Tracy BL, Paxton RJ, et al. Repeated sessions of functional repetitive transcranial magnetic stimulation increases motor cortex excitability and motor control in survivors of stroke. Neuro Rehabil 2013; 33(2): 185-93. George MS, Nahas Z, Molloy M, et al. A controlled trial of daily left prefrontal cortex TMS for treating depression. Biol Psychiatry 2000; 48(10): 962-70. Padberg F, Zwanzger P, Thoma H, et al. Repetitive transcranial magnetic stimulation (rTMS) in pharmacotherapy-refractory major depression: comparative study of fast, slow and sham rTMS. Psychiatry Res 1999; 88(3): 163-71. Berman RM, Narasimhan M, Sanacora G, et al. A randomized clinical trial of repetitive transcranial magnetic stimulation in the treatment of major depression. Biol Psychiatry 2000; 47(4): 332-7. Loo C, Mitchell P, Sachdev P, et al. Double-blind controlled investigation of transcranial magnetic stimulation for the treatment of resistant major depression. Am J Psychiatry 1999; 156(6): 946-8. Hoffman RE, Boutros NN, Hu S, et al. Transcranial magnetic stimulation and auditory hallucinations in schizophrenia. Lancet 2000; 355(9209): 1073-5. Aleman A, Sommer IE, Kahn RS. Efficacy of slow repetitive transcranial magnetic stimulation in the treatment of resistant auditory hallucinations in schizophrenia: a meta-analysis. J Clin Psychiatry 2007; 68(3): 416-21. Vercammen A, Knegtering H, Bruggeman R, et al. Effects of bilateral repetitive transcranial magnetic stimulation on treatment resistant auditory-verbal hallucinations in schizophrenia: a randomized controlled trial. Schizophr Res 2009; 114(1-3): 172-9. Rollnik JD, Huber TJ, Mogk H, et al. High frequency repetitive transcranial magnetic stimulation (rTMS) of the dorsolateral prefrontal cortex in schizophrenic patients. Neuroreport 2000; 11(18): 4013-5. Greenberg BD, George MS, Martin JD, et al. Effect of prefrontal repetitive transcranial magnetic stimulation in obsessivecompulsive disorder: a preliminary study. Am J Psychiatry 1997; 154(6): 867-9. Breiter HC, Rauch SL. Functional MRI and the study of OCD: from symptom provocation to cognitive-behavioral probes of cortico-striatal systems and the amygdala. Neuroimage 1996; 4(3 Pt 3): S127-38. Kang JI, Kim CH, Namkoong K, et al. A randomized controlled study of sequentially applied repetitive transcranial magnetic

[302] [303]

[304] [305] [306]

[307] [308]

stimulation in obsessive-compulsive disorder. J Clin Psychiatry 2009; 70(12): 1645-51. Tergau F, Naumann U, Paulus W, et al. Low-frequency repetitive transcranial magnetic stimulation improves intractable epilepsy. Lancet 1999; 353(9171): 2209. Brasil-Neto JP, de Araujo DP, Teixeira WA, et al. Experimental therapy of epilepsy with transcranial magnetic stimulation: lack of additional benefit with prolonged treatment. Arq Neuropsiquiatr 2004; 62(1): 21-5. Rotenberg A, Bae EH, Muller PA, et al. In-session seizures during low-frequency repetitive transcranial magnetic stimulation in patients with epilepsy. Epilepsy Behav 2009; 16(2): 353-5. Weiss SR, Li XL, Rosen JB, et al. Quenching: inhibition of development and expression of amygdala kindled seizures with low frequency stimulation. Neuroreport 1995; 6(16): 2171-6. Joo EY, Han SJ, Chung SH, et al. Antiepileptic effects of lowfrequency repetitive transcranial magnetic stimulation by different stimulation durations and locations. Clin Neurophysiol 2007; 118(3): 702-8. Mosimann UP, Rihs TA, Engeler J, et al. Mood effects of repetitive transcranial magnetic stimulation of left prefrontal cortex in healthy volunteers. Psychiatry Res 2000; 94(3): 251-6. O'Connor M, Brenninkmeyer C, Morgan A, et al. Relative effects of repetitive transcranial magnetic stimulation and electroconvulsive therapy on mood and memory: a neurocognitive risk-benefit analysis. Cogn Behav Neurol 2003; 16(2): 118-27. George MS, Wassermann EM, Williams WA, et al. Changes in mood and hormone levels after rapid-rate transcranial magnetic stimulation (rTMS) of the prefrontal cortex. J Neuropsychiatry Clin Neurosci 1996; 8(2): 172-80. Pascual-Leone A, Catala MD, Pascual-Leone PA. Lateralized effect of rapid-rate transcranial magnetic stimulation of the prefrontal cortex on mood. Neurology 1996; 46(2): 499-502. Baeken C, Leyman L, De Raedt R, et al. Left and right High Frequency repetitive Transcranial Magnetic Stimulation of the dorsolateral prefrontal cortex does not affect mood in female volunteers. Clin Neurophysiol 2008; 119(3): 568-75. Schutter DJ, van Honk J, d'Alfonso AA, et al. Effects of slow rTMS at the right dorsolateral prefrontal cortex on EEG asymmetry and mood. Neuroreport 2001; 12(3): 445-7. Allan CL, Herrmann LL, Ebmeier KP. Transcranial magnetic stimulation in the management of mood disorders. Neuropsychobiology 2011; 64(3): 163-9. Wassermann EM, Grafman J, Berry C, et al. Use and safety of a new repetitive transcranial magnetic stimulator. Electroencephalogr Clin Neurophysiol 1996; 101(5): 412-7. Kim JY, Chung EJ, Lee WY, et al. Therapeutic effect of repetitive transcranial magnetic stimulation in Parkinson's disease: analysis of [11C] raclopride PET study. Mov Disord 2008; 23(2): 207-11. Mally J, Farkas R, Tothfalusi L, et al. Long-term follow-up study with repetitive transcranial magnetic stimulation (rTMS) in Parkinson's disease. Brain Res Bull 2004; 64(3): 259-63. Shimamoto H, Takasaki K, Shigemori M, et al. Therapeutic effect and mechanism of repetitive transcranial magnetic stimulation in Parkinson's disease. J Neurol 2001; 248 Suppl 3: Iii48-52. Filipovic SR, Rothwell JC, van de Warrenburg BP, et al. Repetitive transcranial magnetic stimulation for levodopa-induced dyskinesias in Parkinson's disease. Mov Disord 2009; 24(2): 246-53. Epstein CM. Seizure or convulsive syncope during 1-Hz rTMS? Clin Neurophysiol 2006; 117(11): 2566-7. Nowak DA, Hoffmann U, Connemann BJ, et al. Epileptic seizure following 1 Hz repetitive transcranial magnetic stimulation. Clin Neurophysiol 2006; 117(7): 1631-3. Wassermann EM. Risk and safety of repetitive transcranial magnetic stimulation: report and suggested guidelines from the International Workshop on the Safety of Repetitive Transcranial Magnetic Stimulation. Electroencephalogr Clin Neurophysiol 1998; 108(1): 1-16. Wassermann EM, Cohen LG, Flitman SS, et al. Seizures in healthy people with repeated safe trains of transcranial magnetic stimuli. Lancet 1996; 347(9004): 825-6. Bernabeu M, Orient F, Tormos JM, et al. Seizure induced by fast repetitive transcranial magnetic stimulation. Clin Neurophysiol 2004; 115(7): 1714-5.

Pe N rs ot on fo al rD U is se tri O bu n tio ly n

[281]

Ghamgosha et al.

[309]

[310] [311]

[312] [313] [314] [315] [316] [317] [318] [319] [320] [321]

[322] [323]

Therapeutic Advances in the Treatment of PD

[325] [326] [327] [328]

[329]

[330]

[331] [332] [333]

[334] [335]

[336]

[337] [338] [339] [340] [341] [342] [343] [344] [345] [346]

Rosa MA, Picarelli H, Teixeira MJ, et al. Accidental seizure with repetitive transcranial magnetic stimulation. J Ect 2006; 22(4): 2656. Grossheinrich N, Rau A, Pogarell O, et al. Theta burst stimulation of the prefrontal cortex: safety and impact on cognition, mood, and resting electroencephalogram. Biol Psychiatry 2009; 65(9): 778-84. Machii K, Cohen D, Ramos-Estebanez C, et al. Safety of rTMS to non-motor cortical areas in healthy participants and patients. Clin Neurophysiol 2006; 117(2): 455-71. Satow T, Mima T, Hara H, et al. Nausea as a complication of lowfrequency repetitive transcranial magnetic stimulation of the posterior fossa. Clin Neurophysiol 2002; 113(9): 1441-3. Janicak PG, O'Reardon JP, Sampson SM, et al. Transcranial magnetic stimulation in the treatment of major depressive disorder: a comprehensive summary of safety experience from acute exposure, extended exposure, and during reintroduction treatment. J Clin Psychiatry 2008; 69(2): 222-32. Anderson BS, Kavanagh K, Borckardt JJ, et al. Decreasing procedural pain over time of left prefrontal rTMS for depression: initial results from the open-label phase of a multi-site trial (OPT-TMS). Brain Stimul 2009; 2(2): 88-92. Xia G, Gajwani P, Muzina DJ, et al. Treatment-emergent mania in unipolar and bipolar depression: focus on repetitive transcranial magnetic stimulation. Int J Neuropsychopharmacol 2008; 11(1): 119-30. Zwanzger P, Ella R, Keck ME, et al. Occurrence of delusions during repetitive transcranial magnetic stimulation (rTMS) in major depression. Biol Psychiatry 2002; 51(7): 602-3. Sakkas P, Mihalopoulou P, Mourtzouhou P, et al. Induction of mania by rTMS: report of two cases. Eur Psychiatry 2003; 18(4): 196-8. Hausmann A, Kramer-Reinstadler K, Lechner-Schoner T, et al. Can bilateral prefrontal repetitive transcranial magnetic stimulation (rTMS) induce mania? A case report. J Clin Psychiatry 2004; 65(11): 1575-6. Ella R, Zwanzger P, Stampfer R, et al. Switch to mania after slow rTMS of the right prefrontal cortex. J Clin Psychiatry 2002; 63(3): 249. Rossi S, Hallett M, Rossini PM, et al. Safety, ethical considerations, and application guidelines for the use of transcranial magnetic stimulation in clinical practice and research. Clin Neurophysiol 2009; 120(12): 2008-39. Benabid AL, Pollak P, Louveau A, et al. Combined (thalamotomy and stimulation) stereotactic surgery of the VIM thalamic nucleus for bilateral Parkinson disease. Appl Neurophysiol 1987; 50(1-6): 344-6. Siegfried J, Lippitz B. Bilateral chronic electrostimulation of ventroposterolateral pallidum: a new therapeutic approach for alleviating all parkinsonian symptoms. Neurosurgery 1994; 35(6): 1126-9. Limousin P, Pollak P, Benazzouz A, et al. Bilateral subthalamic nucleus stimulation for severe Parkinson's disease. Mov Disord 1995; 10(5): 672-4. Patel NK, Heywood P, O'Sullivan K, et al. Unilateral subthalamotomy in the treatment of Parkinson's disease. Brain 2003; 126(Pt 5): 1136-45. Blomstedt P, Fytagoridis A, Tisch S. Deep brain stimulation of the posterior subthalamic area in the treatment of tremor. Acta Neurochir (Wien) 2009; 151(1): 31-6. Blomstedt P, Sandvik U, Fytagoridis A, et al. The posterior subthalamic area in the treatment of movement disorders: past, present, and future. Neurosurgery 2009; 64(6): 1029-38. Volkmann J, Sturm V, Weiss P, et al. Bilateral high-frequency stimulation of the internal globus pallidus in advanced Parkinson's disease. Ann Neurol 1998; 44(6): 953-61. Benabid AL, Pollak P, Gervason C, et al. Long-term suppression of tremor by chronic stimulation of the ventral intermediate thalamic nucleus. Lancet 1991; 337(8738): 403-6. Krack P, Pollak P, Limousin P, et al. Subthalamic nucleus or internal pallidal stimulation in young onset Parkinson's disease. Brain 1998; 121 ( Pt 3): 451-7. Mure H, Hirano S, Tang CC, et al. Parkinson's disease tremorrelated metabolic network: characterization, progression, and treatment effects. Neuroimage 2011; 54(2): 1244-53. Lyons KE, Koller WC, Wilkinson SB, et al. Long term safety and efficacy of unilateral deep brain stimulation of the thalamus for

[347] [348] [349] [350] [351] [352] [353] [354] [355] [356] [357] [358] [359]

[360] [361] [362] [363] [364] [365] [366]

[367] [368] [369]

[370]

223

parkinsonian tremor. J Neurol Neurosurg Psychiat 2001; 71(5): 682-4. Rodriguez MC, Guridi OJ, Alvarez L, et al. The subthalamic nucleus and tremor in Parkinson's disease. Mov Disord 1998; 13 Suppl 3: 111-8. Putzke JD, Wharen RE, Jr., Wszolek ZK, et al. Thalamic deep brain stimulation for tremor-predominant Parkinson's disease. Parkinsonism Relat Disord 2003; 10(2): 81-8. Kumar R, Lozano AM, Sime E, et al. Long-term follow-up of thalamic deep brain stimulation for essential and parkinsonian tremor. Neurology 2003; 61(11): 1601-4. Castrioto A, Lozano AM, Poon YY, et al. Ten-year outcome of subthalamic stimulation in Parkinson disease: a blinded evaluation. Arch Neurol 2011; 68(12): 1550-6. Landi A, Parolin M, Piolti R, et al. Deep brain stimulation for the treatment of Parkinson's disease: the experience of the Neurosurgical Department in Monza. Neurol Sci 2003; 24 Suppl 1: S43-4. Blomstedt P, Sandvik U, Tisch S. Deep brain stimulation in the posterior subthalamic area in the treatment of essential tremor. Mov Disord 2010; 25(10): 1350-6. Krack P, Batir A, Van Blercom N, et al. Five-year follow-up of bilateral stimulation of the subthalamic nucleus in advanced Parkinson's disease. N Engl J Med 2003; 349(20): 1925-34. Rodriguez-Oroz MC, Obeso JA, Lang AE, et al. Bilateral deep brain stimulation in Parkinson's disease: a multicentre study with 4 years follow-up. Brain 2005; 128(Pt 10): 2240-9. Schupbach WM, Chastan N, Welter ML, et al. Stimulation of the subthalamic nucleus in Parkinson's disease: a 5 year follow up. J Neurol Neurosurg Psychiatry 2005; 76(12): 1640-4. Ostergaard K, Aa Sunde N. Evolution of Parkinson's disease during 4 years of bilateral deep brain stimulation of the subthalamic nucleus. Mov Disord 2006; 21(5): 624-31. Volkmann J, Albanese A, Kulisevsky J, et al. Long-term effects of pallidal or subthalamic deep brain stimulation on quality of life in Parkinson's disease. Mov Disord 2009; 24(8): 1154-61. Moro E, Lozano AM, Pollak P, et al. Long-term results of a multicenter study on subthalamic and pallidal stimulation in Parkinson's disease. Mov Disord 2010; 25(5): 578-86. Santos P, Valero R, Arguis MJ, et al. Preoperative adverse events during stereotactic microelectrode-guided deep brain surgery in Parkinson's disease. Rev Esp Anestesiol Reanim 2004; 51(9): 52330. Terao T, Takahashi H, Yokochi F, et al. Hemorrhagic complication of stereotactic surgery in patients with movement disorders. J Neurosurg 2003; 98(6): 1241-6. Binder DK, Rau G, Starr PA. Hemorrhagic complications of microelectrode-guided deep brain stimulation. Stereotact Funct Neurosurg 2003; 80(1-4): 28-31. Constantoyannis C, Berk C, Honey CR, et al. Reducing hardwarerelated complications of deep brain stimulation. Can J Neurol Sci 2005; 32(2): 194-200. Okun MS, Tagliati M, Pourfar M, et al. Management of referred deep brain stimulation failures: a retrospective analysis from 2 movement disorders centers. Arch Neurol 2005; 62(8): 1250-5. Mohammadi A. Induction of neuroplasticity by transcranial direct current stimulation. J Biomed Phys Eng 2016; 6(4): 205-208. Fregni F, Pascual-Leone A. Technology insight: noninvasive brain stimulation in neurology-perspectives on the therapeutic potential of rTMS and tDCS. Nat Clin Pract Neurol 2007; 3(7): 383-93. Aldini G. Essai théorique et expérimental sur le galvanisme; avec une série d'expériences faites en présence des commissaires de l'Institut national de France, et en divers amphithéatres anatomiques de Londres. 1804, Paris: Fournier. Priori A. Brain polarization in humans: a reappraisal of an old tool for prolonged non-invasive modulation of brain excitability. Clin Neurophysiol 2003; 114(4): 589-95. Zago S, Ferrucci R, Fregni F, et al. Bartholow, Sciamanna, Alberti: pioneers in the electrical stimulation of the exposed human cerebral cortex. Neuroscientist 2008; 14(5): 521-8. Kamali AM, Nami M, Yahyavi SS, et al. Transcranial direct current stimulation to assist experienced pistol shooters in gaining even-better performance scores. Cerebellum 2018.: doi: 10.1007/s12311-018-0967-9. [Epub ahead of print]. Fregni F, Boggio PS, Santos MC, et al. Noninvasive cortical stimulation with transcranial direct current stimulation in Parkinson's disease. Mov Disord 2006; 21(10): 1693-702.

Pe N rs ot on fo al rD U is se tri O bu n tio ly n

[324]

Current Gene Therapy, 2018, Vol. 18, No. 4

224 Current Gene Therapy, 2018, Vol. 18, No. 4

[372] [373] [374] [375] [376] [377]

Valentino F, Cosentino G, Brighina F, et al. Transcranial direct current stimulation for treatment of freezing of gait: a cross-over study. Mov Disord 2014; 29(8): 1064-9. Li H, Lei X, Yan T, et al. The temporary and accumulated effects of transcranial direct current stimulation for the treatment of advanced Parkinson's disease monkeys. Sci Rep 2015; 5: 12178. Ferrucci R, Cortese F, Bianchi M, et al. Cerebellar and motor cortical transcranial stimulation decrease levodopa-induced dyskinesias in parkinson's disease. Cerebellum 2016; 15(1): 43-7. Manenti R, Brambilla M, Rosini S, et al. Time up and go task performance improves after transcranial direct current stimulation in patient affected by Parkinson's disease. Neurosci Lett 2014; 580: 74-7. Benninger DH, Lomarev M, Lopez G, et al. Transcranial direct current stimulation for the treatment of Parkinson's disease. J Neurol Neurosurg Psychiat 2010; 81(10): 1105-11. Li Y, Tian X, Qian L, et al. Anodal transcranial direct current stimulation relieves the unilateral bias of a rat model of Parkinson's disease. Conf Proc IEEE Eng Med Biol Soc 2011; 2011: 765-8. Takano Y, Yokawa T, Masuda A, et al. A rat model for measuring the effectiveness of transcranial direct current stimulation using fMRI. Neurosci Lett 2011; 491(1): 40-3.

[378] [379] [380] [381]

[382] [383]

Doruk D, Gray Z, Bravo GL, et al. Effects of tDCS on executive function in Parkinson's disease. Neurosci Lett 2014; 582: 27-31. Boggio PS, Ferrucci R, Rigonatti SP, et al. Effects of transcranial direct current stimulation on working memory in patients with Parkinson's disease. J Neurol Sci 2006; 249(1): 31-8. Tanaka T, Takano Y, Tanaka S, et al. Transcranial direct-current stimulation increases extracellular dopamine levels in the rat striatum. Front Syst Neurosci 2013; 7(6). Winkler C, Reis J, Hoffmann N, et al. Anodal transcranial direct current stimulation enhances survival and integration of dopaminergic cell transplants in a rat parkinson model. Eneuro 2017; 4(5). pii: ENEURO.0063-17. 2017. Poreisz C, Boros K, Antal A, et al. Safety aspects of transcranial direct current stimulation concerning healthy subjects and patients. Brain Res Bull 2007; 72(4-6): 208-14. Nitsche MA, Liebetanz D, Lang N, et al. Safety criteria for transcranial direct current stimulation (tDCS) in humans. Clin Neurophysiol 2003; 114(11): 2220-2.

Pe N rs ot on fo al rD U is se tri O bu n tio ly n

[371]

Ghamgosha et al.