Next Article in Journal
Assessing the Biodegradation of BTEX and Stress Response in a Bio-Permeable Reactive Barrier Using Compound-Specific Isotope Analysis
Previous Article in Journal
Adherence to Combined Healthy Movement Behavior Guidelines among Adolescents: Effects on Cardiometabolic Health Markers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Commentary

New Targets and New Technologies in the Treatment of Parkinson’s Disease: A Narrative Review

1
Department of Neurosurgery, Azienda Ospedaliera Universitaria Pisana (AOUP), University of Pisa, 56100 Pisa, Italy
2
Medicine Faculty, Austral University, Buenos Aires B1406, Argentina
3
Functional Neurosurgery Program, Department of Neurosurgery, San Miguel Arcángel Hospital, Buenos Aires B1406, Argentina
4
Department of Neurosurgery, Hospital Nacional Guillermo Almenara Irigoyen, Lima 07035, Peru
5
Medicine Faculty, Universidad Nacional Mayor de San Marcos, Lima 07035, Peru
*
Author to whom correspondence should be addressed.
Int. J. Environ. Res. Public Health 2022, 19(14), 8799; https://doi.org/10.3390/ijerph19148799
Submission received: 23 May 2022 / Revised: 16 July 2022 / Accepted: 18 July 2022 / Published: 20 July 2022

Abstract

:
Parkinson’s disease (PD) is a progressive neurodegenerative disease, whose main neuropathological finding is pars compacta degeneration due to the accumulation of Lewy bodies and Lewy neurites, and subsequent dopamine depletion. This leads to an increase in the activity of the subthalamic nucleus (STN) and the internal globus pallidus (GPi). Understanding functional anatomy is the key to understanding and developing new targets and new technologies that could potentially improve motor and non-motor symptoms in PD. Currently, the classical targets are insufficient to improve the entire wide spectrum of symptoms in PD (especially non-dopaminergic ones) and none are free of the side effects which are not only associated with the procedure, but with the targets themselves. The objective of this narrative review is to show new targets in DBS surgery as well as new technologies that are under study and have shown promising results to date. The aim is to give an overview of these new targets, as well as their limitations, and describe the current studies in this research field in order to review ongoing research that will probably become effective and routine treatments for PD in the near future.

1. Introduction

Parkinson’s disease (PD) is a chronic, progressive, and debilitating neurodegenerative disease that stands as the second most common neurodegenerative disease and affects about 1% of the population aged above 55 years old [1,2]. PD is the most common cause of “parkinsonism”, a syndrome manifested by rest tremor, rigidity, bradykinesia, and postural instability [3]. These four clinical presentations are commonly present; however, they do not occur long after the disease arises and therefore, they are not included in the Movement Disorder Society (MDS) diagnostic criteria for PD [4,5,6,7,8]. In PD the main neuropathological finding is substantia nigra pars compacta degeneration because of the accumulation of Lewy bodies and Lewy neurites [2,9] and subsequent dopamine depletion [10].
Deep brain stimulation (DBS) is a technique in which one or more multiple-contact electrodes are implanted in specific brain regions. These electrodes are connected to a subcutaneous implantable pulse generator from which the depolarization, re-polarization and conduction patterns of the neurons’ action potential are electrically modulated. Typically, high frequency stimulation (more than 100 Hz) mainly decreases the ipsilateral discharges from the subthalamic nucleus (STN), but there is also a decrease in contralateral STN nucleus firing [11,12,13,14]. Moreover, DBS is a three-dimensionally complex phenomenon, and consequently, a decrease in the action potentials in related nearby areas has been reported [15,16]. Even though there are paradoxical in vitro results published in the literature, it is probably because in vitro samples are disconnected from their afferents, so they do not have the chance for orthodromic stimulation [17,18]. However, aside from the DBS effects, it appears to be transient, with several articles suggesting that DBS as well as new technologies may induce long term changes in neuronal properties [19,20,21]. Neurochemical differences (GABA, glutamate, DA, cGMP, mGLUR, DR1/DR2) and induction of synaptic plasticity have also been observed in the target and in other related deep nuclei, structures, and the cortex itself [16,19,20,21,22,23].
Currently, the FDA (Food and Drug Administration) has approved modulation by DBS in the subthalamic nucleus (STN), the internal globus pallidus (GPi) and the ventralis intermedius nucleus of the thalamus (Vim). Targeting just these classical targets is insufficient for improving the entire wide spectrum of symptoms in PD (especially the non-dopaminergic ones), and none are free of the side effects which are associated not only with the procedure, but with the targets themselves. However, new targets and new technologies that could potentially improve the motor and non-motor symptoms of PD are currently proposed in the literature and are here discussed.

2. Overview of Basal Ganglia Functional Anatomy

The classical model of the basal ganglia includes both the direct and indirect pathways [24,25,26]. Both pathways are modulated by dopamine from the SNc (nigrofugal pathway) which sends projections to the posterior putamen, which in turn sends GABA projections back to the SNc. In addition, the motor cortex, the somatosensory cortex and the STN send glutamatergic fibers to the SNc [24,26,27]. However, there is evidence that the striatofugal pathways (direct and indirect) can no longer be considered dual, due to complex and extensive networks with advanced patterns of bridging collaterals (Figure 1) [28]. This provides straightforward evidence of the overlap and the delicate balance between both pathways [28,29]. Additionally, it is known that there are different functional regions in the striatum and the pallidum for the sensorimotor, associative, and limbic components [30]. For example, the head and the tail of the caudate have mainly associative fibers, while the putamen has mostly sensorimotor connections. Additionally, the ventral striatum has predominantly associative fibers and is also related to the limbic system [30,31]. However, in the end it is difficult to draw a boundary line between the dorsal and the ventral striatum, as well as between the functionality of each one [30].
The pallidofugal pathway is related to symptoms such as tremor and stiffness [32] and it is also known that 90% of cells that make up the GPi are motor neurons which connect to the PPN, ventral and CM/PF nuclei of the thalamus; however, there are 10% of limbic neurons that reach the lateral habenula [25]. Traditionally, it was believed that in the pallidofugal pathway the bipolar neurons at the lateral GPi would send their axons to the thalamus through the lenticular fasciculus and the ansa lenticularis, while neurons at the medial GPi would only send their axons through the lenticular fasciculus, and the GPe would send its efferents through the subthalamic fasciculus to reach the STN (Figure 2a) [25,33,34]. Nevertheless, there is evidence that the pallidofugal pathway fibers from the dorsal part of the GPi travel through the lenticular fasciculus, then pass between the STN and the ZI to finally reach the ventral part of the thalamus, while the fibers from the intermediate part of the GPi travel through the pallidum subthalamic tract and reach the lateral STN side [25,33,35]. Finally, the axons from the ventral part of the GPi travel through the ansa lenticularis then connect to the SN and continue to the brainstem through the fasciculus Q of Sano or the pallido tegmentalis, while other fibers of the ansa lenticularis turn dorsally to reach the thalamus through the thalamic fasciculus (Figure 2b) [35]. In the case of the GPe, there is a description of a direct projection to the frontal regions of the cerebral cortex, the caudate and the putamen [34,36].
With regards to the cerebral cortex, its connections with the basal ganglia are also well known (Corticofugal pathway). For instance, the prefrontal cortex has connections with the caudate’s head; however, the dorsolateral prefrontal cortex and the pre-supplementary motor area are related to the anterior putamen [2,26]. Meanwhile, the limbic cortical areas that are connected to the ventral striatum and the motor cortex by the corticostriatal pathway, reach the posterior putamen [2,26,30]. The study of these pathways resulted in the identification of the glutamatergic hyperdirect cortico-subthalamo-pallidal circuit which plays a very important role in DBS procedures [24,37]. Furthermore, functional images have provided evidence of cerebellar activation in a wide variety of motor and non-motor processes [38,39,40,41]. It has been proved that a lot of cortical areas and basal ganglia nuclei, which are targets of cerebellar output, also project back onto the cerebellum (Figure 3) [40,42,43]. STN DBS may not only normalize the functional activity of the associative and limbic cerebellar cortices and decrease the activity of Purkinje cells with consequent disinhibition of the cerebellar nuclei, but could also improve some non-motor symptoms [40,44,45]. There is also evidence that STN DBS in dystonia, Tourette syndrome and PD results in decreased cerebellar hypermetabolism and symptom improvement [40]. Research in functional neuroanatomy has allowed critical advances in the development of new devices as well as the possibility of new targets.

3. New DBS Targets

3.1. Pedunculopontine Nucleus

The pontine peduncle nucleus (PPN) is a motor center of the trunk that is related to the initiation and modulation of the gait and receives afferents from the cortex, the central core, the trunk, and the medulla [46,47]. The PPN consists of the Cholinergic Pars Compacta (PPNc) and the Pars Dissipatus (glutamatergic and cholinergic) [47] and it is located medially to the middle lemniscus and laterally to the decussation of the superior cerebellar peduncle. Its rostral portion is posterior to the SN and the RRF, while its caudal pole meets the reticular formation, the cuneiform nucleus, and the locus coeruleus [48]. It has been observed that low-frequency stimulation of the PPN (bilateral or unilateral) inhibits the GABAergic neurons of this nucleus, which allows their activation and increases movement [47,49,50]. It is important to note that this nucleus may be involved in PD. However, the stimulation of this area may compromise adjacent structures such as the caudal portion of the mesencephalic reticular formation [47,48,49,51]. In many cases, DBS of the PPN has had good results on different aspects of the gait and appears to be a safe target [51]. Nevertheless, this effect is not uniform in all studies, nor in all patients [52,53,54].

3.2. Zona Incerta

The caudal zona incerta is a known target and is part of a region called the posterior subthalamic area, which is located posteromedial to the STN, inferior to the thalamus and lateral to the red nucleus [9]. It has been observed in studies carried out in patients with STN-DBS, that stimulation of this area and the adjacent ones (pallidothalamic fibers) improves the results, which contrasts with patients where only the STN is stimulated, thus developing interest in the surrounding structures (the zona incerta) as a possible therapeutic target since it is a unique GABAergic link between the basal ganglia output nuclei, the cerebello-thalamo-cortical loop and the brainstem nuclei [55,56,57,58].
Due to their interconnections with the basal ganglia and the cerebellum (Figure 1), stimulation of the zona incerta (incerto-thalamic fibers) and the bundles that cross this area (the nigrothalamic fibers, the incerto-pallidal bundle, the pallidal-thalamic pathway and the cerebellar-thalamic tract) improves, importantly, symptoms such as essential voice tremor compared to STN stimulation [46,59], and also results in decreases in bradykinesia, rigidity, ataxia and abnormal muscle activities [60,61]. Moreover, stimulating the ZI and adjacent areas decreases the required dose, and reduces the adverse effects, of dopaminergic medication [62]. Other points in favor of this target are that it is relatively easy for neurosurgeons to leave the deep ventral DBS contact in the caudal ZI, and that long term studies have shown that the ZI does not develop tolerance; however, Vim DBS did provide better outcomes [61,63]. However, it is necessary to highlight that it can cause or exacerbate dyskinesia, dysarthria, and can also alter pleasure-seeking behavior [60,64]. There are limitations to these findings with regards to it being a new target so it is a procedure rarely performed and with few studies on the subject, as such, we cannot define conclusions [61,63].

3.3. Thalamic CM/Pf Complex

The thalamus as a target (Vim) for DBS was the first to be studied and applied in clinical practice and demonstrated adequate control of the tremor at rest [65]. The centromedian/parafascicular complex (CM/Pf) is located in the posterior intralaminar thalamus [66,67]. It is an important center for sensory, limbic and motor processing association due to its extensive connections with the cortex, the basal ganglia and the brainstem [66]. It has been observed in patients whose dyskinesia improves that the thalamic electrode is closer to the CM/Pf than to the Vim [67,68]. Similarly, clinical improvement has been highlighted in tremor and freezing of gait, motivating interest in its role in DBS surgery [69,70]. Although it does not affect extrapyramidal symptoms with the same intensity as DBS of the STN, it seems to be a better target than the Vim and it opens up the possibility of a synergistic and complementary effect with other targets for patients with difficult-to-control symptoms [67,69,70,71].

3.4. Substantia Nigra Pars Reticulata (SNr)

The midbrain locomotor region is a functional territory that appears to be comprised of both the cuneiform nucleus and part of the PPN and is directly influenced by GABAergic nigro-tegmental projections from the SNr in addition to the supplemental motor cortex [72,73]. This relationship of the SNr with the ponto-mesencephalic and spinal structures motivated experimental studies where it was observed that both its unilateral and bilateral stimulation (and in many of these cases combining its effect with the stimulation of the STN), could favor the improvement of axial symptoms related to the onset of gait in patients with PD [72,74]. Furthermore, an improvement has also been noted in the axial symptom subsection of the UPDRS scale, thus constituting a potential tool for the treatment of postural instability and freezing of the gait [73,75,76]. However, psychiatric side effects (acute depression, hypomania and mania) have been reported in a variable way and are possibly related to its limbic connections [73].

3.5. Multitarget

Multitarget therapy is a novel and interesting option that involves more than one anatomical objective for patients in whom the most common targets such as STN or GPi, by themselves, are insufficient for controlling the symptoms of the disease, mostly, non-dopaminergic symptoms such as gait, balance impairment and cognitive decline [46,77,78]. As detailed in the previous sections, the use of new targets such as the PPN, the ZI, the SNr and the thalamus for DBS is under investigation [54]. These options are open to different therapeutic scenarios for the management of complex cases. If their effectiveness is demonstrated, it is most likely that they will not replace the classical targets but will be used as adjuvants in selected cases [79]. Similarly, it is possible to combine the stimulation of two classical targets such as the STN and the GPi to obtain better results [80]. The development of multitarget therapy is based on the improvement of surgical times, technological progress and the clinical benefits associated with biochemical changes in areas other than the STN for the legitimate control of clinical signs that cannot be controlled with classical DBS targets [70,80]. Candidates for multitarget therapy are patients with involuntary movements who are reluctant to decrease their dose of drugs; patients with severe tremor whose response to STN-DBS progressively decreases; patients freezing in ON periods; and even patients with mild cognitive impairment whose eligibility for STN-DBS may be in doubt [80].
One of the advantages of GPi/STN-DBS in PD is that it offers greater control of neuropsychiatric symptoms. [46,70] Currently, a GPi/STN-DBS strategy is more than an eligible approach for those cases where there is profound and medically refractory functional impairment [78,81]. GPi/STN-DBS has shown favorable outcomes; however, its effect has been seen to decrease over time. For this reason, more evidence is needed to fully understand the advantages and limitations of this dual stimulation [78]. Treating multiple targets with DBS is limited by its higher costs, the prolonged inconvenience of surgery due to its slower result, greater probability of complications, and the requirement of multidisciplinary and highly specialized healthcare staff for peri- and post-operative management [70,77]. This review is susceptible to changes through time, as technology and science become more advanced and new research will be reported in the near future.

4. Non-Invasive and Minimally Invasive Treatment

4.1. Endovascular Approach

The use of endovascular devices (with cable or without cable) with the ability to record brain activity and stimulate adjacent tissue (STENTrodes) has emerged as an alternative among mechanisms of deep stimulation with minimal invasion for the treatment of PD, essential tremor, epilepsy, severe paralysis and many other diseases [82,83]. There is a wide variety of catheters, guide catheters, microcatheters and micro guides, among others, which facilitate the release of the devices in the desired location [82,84]. The endovascular routes for the release of the device can be by a transarterial or transvenous route [83]. However, the anatomical variations of the vascular structures are a limitation for the adequate positioning of the device, since precise planning for the identification of an accessible vascular structure (vessels with a diameter greater than 1 mm are more accessible) closest to the target is essential. Angiography and MRI protocols have taken on a greater relevance for the elaboration of probabilistic maps in order to improve the accuracy of the device placement [83,85,86]. Only some structures such as the nucleus accumbens, fornix, subcallosal cortex of the cingulum, dentato-rubro-thalamic tract, subthalamus and pontine peduncle nucleus, among others, are susceptible to stimulation by this route due to their vascular proximity [83,87] (Figure 4). It should be noted that the possibility of performing the procedure with the patient awake allows for evaluation of the procedure and observing whether the result is the desired one [88]. Although endovascular therapy has an encouraging future due to new and continuous advances, it must be considered that there are potential side effects depending on the location of the device, ranging from psychiatric disorders (due to proximity to the limbic system) to autonomic problems (due to proximity to the hypothalamus), and also those of the procedure such as intracranial hemorrhage and arterial or venous thrombosis by the intraluminal device (with SHIELD technology the thrombogenicity of the devices is reduced) [82,83,87,89].

4.2. Non-Invasive Transcranial Stimulation

Non-invasive transcranial stimulation is based on beta oscillatory activity (14–35 Hz anti-kinetic activity) that occurs at the level of the motor cortex and its connections with the basal ganglia (predominantly with the subthalamus and the GPe) [90,91,92]. These connections are exacerbated (excessive beta synchronization) in patients with PD [91,92,93,94]. This abnormal activity occurs mainly during involuntary tonic contractions and decreases with voluntary motor activity (beta desynchronization), as well as with pharmacological and surgical treatment [91,92,93,94].
Currently, there are two techniques known as “Non-invasive transcranial stimulation”: The first one, Transcranial Current Magnetic Stimulation (TMS) and the second one, Transcranial Current Stimulation (tCS) [95,96]. Both have proved to have a good efficacy rate in the modulation of cerebral activity [95,96,97]. One of the advantages of this non-invasive technology over DBS is that it can target cortical and subcortical structures; therefore, it can reach remote locations from the stimulation site [96,98]. It should also be mentioned that DBS is an invasive procedure, consequently, it goes hand in hand with some complications, such as infection, limited duration of electrical components, neural immune system reactions, and the need for periodic battery replacement [99,100,101,102]. One of the limitations of TMS is that this device is large and heavy, whereas tDCS is light and small, which makes it more feasible since it is a treatment that does not require hospitalization and the sessions can be performed at the patient’s home [97,103]. Cost is another big difference, with TMS systems costing between $20,000 and $100,000, while tDCS devices have prices ranging from $400 to $10,000 [95,97,104].

4.2.1. Transcranial Magnetic Stimulation (TMS)

TMS is a well-tolerated and painless brain stimulation technique in which a strong and rapidly changing electromagnetic field is generated [95,97,103,105]. This electromagnetic field induces strong and short-lived electrical currents, which initiate action potentials in both the cortex and subcortical white matter [96,97,105]. The application of repetitive transcranial magnetic stimulation (rTMS) is able to induce lasting changes in cortical excitability, which represents a promising tool against neuropsychiatric alterations and motor symptoms in PD [98,104,106]. Unlike TMS, rTMS appears to be more effective in stimulating the association cortex [107]. Presently, TMS has a larger number of published studies demonstrating its safety in terms of effects on brain anatomy and biochemistry, than does tDCS [97,105]. Other studies being conducted, which link rTMS with an improvement in non-motor symptoms in PD such as cognitive dysfunction, speech difficulty and depression, show mixed results [105,107]. However, its greatest risk is the induction of seizures [108]. Preclinical data on the safety of TMS is still very limited, and rTMS protocols are needed to use this technology as a routine treatment for PD [95,98].

4.2.2. Transcranial Current Stimulation—Transcranial Direct Current Stimulation (tDCS)

TCS has re-emerged as a form of non-invasive modulation of spontaneous neuronal activity by applying a weak electrical current (1–2 mA) through the placement of two or more electrodes on the scalp, allowing regulation of neuronal membrane potentials (changes in discharge probability), changes in neurotransmitters, glia, and micro vessels [109]. In other words, it provides neuronal plasticity, shown by the generation of long-term synaptic potentiation (objectified by the increase in brain derived neurotrophic factor (BDNF) secretion and TrkB activation) when combined with repetitive low-frequency synaptic activation [110]. Transcranial direct current stimulation (tDCS) affects the motor symptoms of the disease, with the most prominent results related to rehabilitation. However, its usefulness is limited due to its weak effects, high variability, and lack of consensus in protocols, with medication status being a key cofounder in determining the level of efficacy [111].

4.2.3. Transcranial Current Stimulation—Transcranial Stimulation with Alternating Current (tACS)

Recent innovations in transcranial alternating current stimulation (tACS) offer new areas of research [111]. This is a variant of direct cranial stimulation and consists of the interference of sinus waves for the modulation of brain oscillations (physiological brain activity in an area) [109,111]. It has been observed that tACS has a different effect on healthy people compared to those with PD due to its effect on beta synchronization [92,112]. tACS seems to produce an improvement in motor symptoms in people with PD, which is further enhanced when associated with a closed loop system [92]. Furthermore, when this device stimulates the cerebellum, it also results in the modulation of parkinsonian tremor [113]. In addition to this, there is evidence that, like tDCS, it not only modulates brain activity, but also seems to generate cortical neuronal plasticity [114,115,116].

4.3. Non-Invasive Focused Ultrasound

Magnetic resonance-guided focused ultrasound (MRgFUS) is a novel, non-invasive procedure for symptomatic treatment of PD [117,118]. Two types of focused ultrasound therapy are recognized. The first one is called non-ablative: this technology has potential uses in pain neuromodulation, epilepsy, and is also seen as a new strategy to allow the passage of medication through the BBB, for instance, biological-genetic treatment in PD [117,119]. The second one is ablative ultrasound, in which two varieties are differentiated according to the frequency used. High intensity focused ultrasound (HIFU) produces thermal ablation in small brain targeted areas (therapeutic sonication) [118,120]. This technology enables accurate targeting by using a real-time MRI for morphological and thermal monitoring [120,121]. On the other hand, ablation at low frequencies is due to a fast change in the targeted tissue pressure, forming gas–vapor filled cavities. These cavitation bubbles oscillate at large amplitudes and exert shear stresses on the surrounding tissue, causing the mechanical cell membranes to tear (histotripsy) [119,122,123].
The use of HIFU in the Vim nucleus has been approved by the FDA for the treatment of tremor in patients with PD [117,120,124]. Candidate patients for Vim ablation are those with very asymmetric symptoms, who are not able to have or do not want surgery or implants, but does not including patients with refractory tremor. Exclusion criteria are severe bilateral axial symptoms, severe dyskinesia, cognitive impairment, or psychiatric illness [117]. Although unilateral ablation of the Vim seems to be effective, it has transitory side effects such as weakness, gait, and speech disturbances [120]. The STN is a potential target for MRgFUS; however, due to few studies being available, it cannot be considered as strong evidence. This procedure has a high rate of adverse effects, with gait instability being the most frequent; however, all are mostly transient [117,120,125]. Similarly, the results of GPi ablation in patients with PD is little but encouraging due to the effects its ablation has on dyskinesias, as well as the improvement of PD symptoms in off-L-dopa state patients [117,126]. Furthermore, challenges lie ahead in treating this target due to its proximity to the optic nerve and being able to correctly target the ultrasound rays on the treatment area [118,127].
The advantages of MRgFUS over DBS are that this non-invasive technology does not carry any risk of infection or hardware failure, and post-operative programming is not required as the effect is achieved immediately at the end of the procedure [120,128]. What is more, this procedure does not include any kind of brain penetration, therefore, it does not have the complications associated with foreign object implantation [120,129,130]. Gamma knife use also effectively suppresses tremor but is limited by the latent effects of radiation and the inability to target intraoperatively [130]. Nevertheless, some disadvantages of MRgFUS lie in its local effects, such as headache, scalp burns, and bone necrosis [123].

5. Regenerative Medicine in Parkinson Disease

Currently, none of the pharmacological or surgical treatments can cure or modify the neurodegeneration process that occurs in PD [131,132]. Furthermore, it is known that the current management of PD is insufficient and costs billions to the world economy. In the USA alone, the cost of PD is 35 billion annually and it is estimated that by stopping its progression, the economic benefit would be almost 450,000 dollars per patient [133]. Regenerative medicine has been progressively developed in recent decades with the aim of reconstructing and restoring the functionality of highly sophisticated neural circuits (connectome) that are lost in PD [131,132,134,135]. Technological progress in cell engineering, cell programming, immunomodulation, tissue engineering, biomaterials, gene therapy, and stem cells has allowed the development of several studies that project a promising future in this type of biological therapy [136,137,138,139,140].

5.1. Gene, Cell, and Tissue Regenerative Therapies

Autologous or heterologous cell transplantation with neural differentiation potential has shown satisfactory results in animals and humans [136,141,142]. Different cell types can be divided into these groups: fetal mesencephalic cells, adrenal medullary cells, carotid body cells, sympathetic ganglia cells, and retinal pigmentary epithelial cells [142,143]. Human fetal mesencephalic tissue grafts are the ones with the most evidence and the greatest success (documented improvement of motor symptoms) to date [144,145,146]. The objectives of this type of therapy are the survival of dopaminergic cells, afferent and efferent synaptic integration, and adequate and controlled dopamine release, in addition to the clinical improvement of patients [136,141,147]. There are reports of patients in whom the presence of transplanted cells has been demonstrated up to 24 years later; however, their potential susceptibility to degeneration is still not clear [145,148]. The problems with the clinical studies involving fetal mesencephalic cells include: the heterogeneity of the cells, in most cases the transplant has not been performed in the SNc region but in the striatum, there is a well-known limitation in the availability of fetal tissue, cells do not always differentiate as expected, it has been observed that the grafts are not made up purely of dopaminergic cells, this kind of procedure often requires immunosuppression, and there are underlying ethical considerations [142,149,150,151]. Another risk in the use of stem cells is the neoformation of tumors and the development of graft-induced dyskinesia (GID) [152,153,154].
There are other sources of cells with potential dopaminergic differentiation such as embryonic stem cells (ESC), induced pluripotential stem cells (iPSC), neural precursor cells (NPC), mesenchymal stem cells (MSC), and direct neuronal reprogramming [152,155,156]. ESCs come from early-stage embryos, which implies ethical problems, limited availability, and they have a significant risk of tumor development [152]. Knowledge of cell de-differentiation techniques as well as advances in neurodevelopment have allowed the generation of iPSCs with greater similarity to dopaminergic neurons and with a lower oncological risk [156,157]. NSC can be induced from iPSC and because of their limited differentiation capability they have less oncogenic likelihood [158]. On the other hand, MSCs have a more limiting potential for neural differentiation but have an immunomodulatory effect that counteracts neurodegeneration [159]. Problems with earlier grafts have generated an increased interest in neuronal reprogramming, in which virus vectors, microRNAs, or transcription factors can be used to alter the gene expression in astrocytes in order to generate dopaminergic neurons [138,140].
Despite advances in cell restoration, there is a gap to recover the lost functionality, that is, the development of new and appropriate neuronal circuits which maintain the dopaminergic function [139,160,161,162,163]. Axonal growth over long distances, its directionality, and the reinnervation of the correct target continue to be a matter of investigation [160,164,165,166,167]. Nowadays, chemoattractant factors are being used in the striatum to promote the restructuring of the nigrostriatal pathway [165]. Other strategies include the use of embryonic striatal tissue, renal tissue, kainate injections, and overexpression of GDNF (glial cell line-derived neurotrophic factor) in the striatum by virus vectors [168,169]. The development of new micro biomaterials has allowed the creation of scaffolds for grafts, that not only protect and promote the graft growth but can also be implanted in the form of neural networks that stimulate the nigrostriatal pathway and favor its restoration [139,160,170]. Finally, there is a potential risk that any cell or tissue engineering strategy may succumb to the underlying pathology and degenerate in a similar way to the original tissue, so developing new grafts resistant to different stress factors is a new objective [151,171,172].

5.2. Optogenetic Therapy

The lack of control over the delimitation of the stimulated area can lead to undesirable results [173,174]. Due to this problem, optogenetic therapy has emerged as an interesting option to avoid side effects due to the stimulation of structures adjacent to the target [175]. This therapy consists of the use of genetically altered viruses with the aim of modifying neuronal genes (the gene for an ultra-fast opsin called Chronos is packaged in the adeno-associated virus type V with a CaMKII promoter for a local effect on the subthalamic nucleus) and making them susceptible to excitation or inhibition by light, so that the desired effect is generated in a more controlled way in each area [171,172,173,174,175]. Optogenetic therapy opens a way for us to understand synaptic and circuit properties on a mechanistic level [176,177]. Favorable results in several animal studies lead us to believe that this therapy could be an important research tool for future DBS-based therapies, leading to symptoms being resolved, and not masking them [174,178]. However, to date this technique is not yet feasible in humans and suffers from other limitations such as light stimulation parameters (duration, frequency, and intensity of stimulation) which must be managed carefully to avoid problems such as depolarization blocks or rebound excitation [177,179,180,181]. We hope that one day this could be a viable treatment for movement disorders.

6. Discussion and Conclusions

Research into functional anatomy as well as clinical trials have allowed a better understanding of the basis, benefits, and limitations of DBS procedures [24,26,38]. The most common scope of studies in PD is directed towards motor symptoms [132]. However, non-motor symptoms are as relevant as these but less studied and, due to their great impact on quality of life, they require more intensive research [73,132]. Currently, progress in the development of devices is reflected in increasingly personalized treatments and the approach of new targets [46,77]. Although most of these targets require more evidence for their widespread use, they open the possibility of being associated with classical targets to achieve better symptom control [79,80]. One of the main disadvantages of DBS is the high cost involved while newer devices appear to be cheaper [92,104,133]. In addition, several of these new devices offer the advantage of being minimally invasive or non-invasive at all [85,87,105].
Accelerated advances in biological therapy, gene therapy, cell engineering, tissue engineering, and biomaterials have made it possible to propose new therapeutic options with the potential to modify the course of the disease [134,140,160]. Even though several studies have shown encouraging long-term results with the use of embryonic or fetal human cells, the limitations and risks of this type of graft have motivated the development of techniques for the induction of new cells with dopaminergic potential [151,156,157]. Nevertheless, we must consider that both the management of fetal or embryonic tissue, as well as genetic manipulation at the cellular or viral level, entail ethical considerations and risks [152,155,169]. The development and use of biomaterials in the peripheral nervous system is currently quite widespread; however, the central nervous system constitutes a more complex area for their use. The accuracy in the design, its microstructure that tries to reproduce the neural circuits, and the advantages shown in the studies are encouraging [164,169]. Most current studies are focused on technological advances, but this should not be separated from the study of functional micro neuroanatomy, since the integration and deepening of both will allow for better results.
There is parallel improvement in regenerative and non-regenerative treatment. Non-regenerative treatment seeks to improve the performance, invasiveness issues, and duration of its devices [46,77,96]. However, it has not been shown to affect the disease progression [118]. Merely stopping its progression would generate a major impact, but current initiatives go beyond that and seek to regenerate the tissue and lost connections, that is, to cure it [133,160]. As described, the new therapeutic options also have limitations and problems, but under the same logic of multitarget treatment, with the wide range of current options, studies of multimodal or hybrid management in PD should be considered. We believe that the treatment of PD will no longer be symptomatic related in due time and that the future directions of surgical management of PD are towards regenerative neurosurgery. We have described the current studies and ongoing research in PD in order to review the treatments that may become effective and safe in the near future for the treatment of PD.

Author Contributions

Conceptualization, J.L., N.A., P.G. and N.M.; methodology, N.A. and N.M.; validation, J.L., N.A. and A.E.; formal analysis, J.L., N.A. and P.G.; writing—original draft preparation, J.L., N.A., P.G. and N.M.; writing—review and editing, J.L., N.A., A.E. and N.M.; visualization, J.L.; supervision, J.L., N.A. and N.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ascherio, A.; Schwarzschild, M.A. The epidemiology of Parkinson’s disease: Risk factors and prevention. Lancet Neurol. 2016, 15, 1257–1272. [Google Scholar] [CrossRef]
  2. Foffani, G.; Obeso, J.A. A Cortical Pathogenic Theory of Parkinson’s Disease. Neuron 2018, 99, 1116–1128. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Obeso, J.A.; Rodriguez-Oroz, M.C.; Goetz, C.G.; Marin, C.; Kordower, J.H.; Rodriguez, M.; Hirsch, E.C.; Farrer, M.; Schapira, A.H.; Halliday, G. Missing pieces in the Parkinson’s disease puzzle. Nat. Med. 2010, 16, 653–661. [Google Scholar] [CrossRef]
  4. Gelb, D.J.; Oliver, E.; Gilman, S. Diagnostic criteria for Parkinson disease. Arch. Neurol. 1999, 56, 33–39. [Google Scholar] [CrossRef] [PubMed]
  5. Hughes, A.J.; Daniel, S.E.; Kilford, L.; Lees, A.J. Accuracy of clinical diagnosis of idiopathic Parkinson’s disease: A clinico-pathological study of 100 cases. J. Neurol. Neurosurg. Psychiatry 1992, 55, 181–184. [Google Scholar] [CrossRef] [Green Version]
  6. Langston, J.W.; Widner, H.; Goetz, C.G.; Brooks, D.; Fahn, S.; Freeman, T.; Watts, R. Core assessment program for intracerebral transplantations (CAPIT). Mov. Disord. 1992, 7, 2–13. [Google Scholar] [CrossRef]
  7. Ward, C.D.; Gibb, W.R. Research diagnostic criteria for Parkinson’s disease. Adv. Neurol. 1990, 53, 245–249. [Google Scholar]
  8. Postuma, R.B.; Berg, D.; Stern, M.; Poewe, W.; Olanow, C.W.; Oertel, W.; Obeso, J.; Marek, K.; Litvan, I.; Lang, A.E.; et al. MDS clinical diagnostic criteria for Parkinson’s disease. Mov. Disord. 2015, 30, 1591–1601. [Google Scholar] [CrossRef]
  9. Kalia, L.V.; Lang, A.E. Parkinson’s disease. Lancet. 2015, 386, 896–912. [Google Scholar] [CrossRef]
  10. Hornykiewicz, O. The discovery of dopamine deficiency in the parkinsonian brain. J. Neural. Transm. Suppl. 2006, 70, 9–15. [Google Scholar]
  11. Burbaud, P.; Gross, C.; Bioulac, B. Effect of subthalamic high frequency stimulation on substantia nigra pars reticulata and globus pallidus neurons in normal rats. J. Physiol. Paris 1994, 88, 359–361. [Google Scholar] [CrossRef]
  12. Filali, M.; Hutchison, W.D.; Palter, V.N.; Lozano, A.M.; Dostrovsky, J.O. Stimulation-induced inhibition of neuronal firing in human subthalamic nucleus. Exp. Brain Res. 2004, 156, 274–281. [Google Scholar] [CrossRef] [PubMed]
  13. Welter, M.-L.; Houeto, J.-L.; Bonnet, A.-M.; Bejjani, P.-B.; Mesnage, V.; Dormont, D.; Navarro, S.; Cornu, P.; Agid, Y.; Pidoux, B. Effects of high-frequency stimulation on subthalamic neuronal activity in parkinsonian patients. Arch. Neurol. 2004, 61, 89–96. [Google Scholar] [CrossRef] [PubMed]
  14. Shi, L.H.; Luo, F.; Woodward, D.J.; Chang, J.Y. Basal ganglia neural responses during behaviorally effective deep brain stimulation of the subthalamic nucleus in rats performing a treadmill locomotion test. Synapse 2006, 59, 445–457. [Google Scholar] [CrossRef]
  15. Tai, C.-H.; Boraud, T.; Bezard, E.; Bioulac, B.; Gross, C.; Benazzouz, A. Electrophysiological and metabolic evidence that high-frequency stimulation of the subthalamic nucleus bridles neuronal activity in the subthalamic nucleus and the substantia nigra reticulata. FASEB J. 2003, 17, 1820–1830. [Google Scholar] [CrossRef] [Green Version]
  16. McIntyre, C.C.; Anderson, R.W. Deep brain stimulation mechanisms: The control of network activity via neurochemistry modulation. J. Neurochem. 2016, 139, 338–345. [Google Scholar] [CrossRef] [Green Version]
  17. Bosch, C.; Degos, B.; Deniau, J.M.; Venance, L. Subthalamic nucleus high-frequency stimulation generates a concomitant synaptic excitation–inhibition in substantia nigra pars reticulata. J. Physiol. 2011, 589, 4189–4207. [Google Scholar] [CrossRef]
  18. Johnson, M.D.; Miocinovic, S.; McIntyre, C.C.; Vitek, J.L. Mechanisms and targets of deep brain stimulation in movement disorders. Neurotherapeutics 2008, 5, 294–308. [Google Scholar] [CrossRef] [Green Version]
  19. Creed, M.; Pascoli, V.J.; Lüscher, C. Addiction therapy. Refining deep brain stimulation to emulate optogenetic treatment of synaptic pathology. Science 2015, 347, 659–664. [Google Scholar] [CrossRef]
  20. Hescham, S.; Temel, Y.; Schipper, S.; Lagiere, M.; Schönfeld, L.M.; Blokland, A.; Jahanshahi, A. Fornix deep brain stimulation induced long-term spatial memory independent of hippocampal neurogenesis. Brain Struct. Funct. 2017, 222, 1069–1075. [Google Scholar] [CrossRef] [Green Version]
  21. Ni, Z.; Kim, S.J.; Phielipp, N.; Ghosh, S.; Udupa, K.; Gunraj, C.A.; Saha, U.; Hodaie, M.; Kalia, S.K.; Lozano, A.M. Pallidal deep brain stimulation modulates cortical excitability and plasticity. Ann. Neurol. 2018, 83, 352–362. [Google Scholar] [CrossRef]
  22. Cheng, Y.; Xie, X.; Lu, J.; Gangal, H.; Wang, W.; Melo, S.; Wang, X.; Jerger, J.; Woodson, K.; Garr, E. Optogenetic induction of orbitostriatal long-term potentiation in the dorsomedial striatum elicits a persistent reduction of alcohol-seeking behavior in rats. Neuropharmacology 2021, 191, 108560. [Google Scholar] [CrossRef] [PubMed]
  23. Stefani, A.; Trendafilov, V.; Liguori, C.; Fedele, E.; Galati, S. Subthalamic nucleus deep brain stimulation on motor-symptoms of Parkinson’s disease: Focus on neurochemistry. Prog. Neurobiol. 2017, 151, 157–174. [Google Scholar] [CrossRef] [PubMed]
  24. Simonyan, K. Recent advances in understanding the role of the basal ganglia. F1000Research 2019, 8, F1000 Faculty Rev-122. [Google Scholar] [CrossRef]
  25. Parent, M.; Parent, A. The pallidofugal motor fiber system in primates. Parkinsonism Relat. Disord. 2004, 10, 203–211. [Google Scholar] [CrossRef]
  26. Obeso, J.A.; Stamelou, M.; Goetz, C.G.; Poewe, W.; Lang, A.E.; Weintraub, D.; Burn, D.; Halliday, G.M.; Bezard, E.; Przedborski, S.J.M.D. Past, present, and future of Parkinson’s disease: A special essay on the 200th Anniversary of the Shaking Palsy. Mov. Disord. 2017, 32, 1264–1310. [Google Scholar] [CrossRef] [PubMed]
  27. Trutti, A.C.; Mulder, M.J.; Hommel, B.; Forstmann, B.U. Functional neuroanatomical review of the ventral tegmental area. NeuroImage 2019, 191, 258–268. [Google Scholar] [CrossRef] [Green Version]
  28. Wu, Y.; Richard, S.; Parent, A. The organization of the striatal output system: A single-cell juxtacellular labeling study in the rat. Neurosci. Res. 2000, 38, 49–62. [Google Scholar] [CrossRef]
  29. Simonyan, K.; Cho, H.; Hamzehei Sichani, A.; Rubien-Thomas, E.; Hallett, M. The direct basal ganglia pathway is hyperfunctional in focal dystonia. Brain 2017, 140, 3179–3190. [Google Scholar] [CrossRef] [Green Version]
  30. Voorn, P.; Vanderschuren, L.J.; Groenewegen, H.J.; Robbins, T.W.; Pennartz, C.M. Putting a spin on the dorsal–ventral divide of the striatum. Trends Neurosci. 2004, 27, 468–474. [Google Scholar] [CrossRef]
  31. Grabli, D.; McCairn, K.; Hirsch, E.C.; Agid, Y.; Féger, J.; François, C.; Tremblay, L. Behavioural disorders induced by external globus pallidus dysfunction in primates: I. Behavioural study. Brain 2004, 127, 2039–2054. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Geibl, F.F.; Henrich, M.T.; Oertel, W.H. Mesencephalic and extramesencephalic dopaminergic systems in Parkinson’s disease. J. Neural. Transm. 2019, 126, 377–396. [Google Scholar] [CrossRef]
  33. Parent, A.; Cossette, M.; Lévesque, M. Anatomical Considerations in Basal Ganglia Surgery. In Movement Disorder Surgery; Karger Publishers: Basel, Switzerland, 2000; Volume 15, pp. 21–30. [Google Scholar]
  34. Sato, F.; Lavallée, P.; Lévesque, M.; Parent, A. Single-axon tracing study of neurons of the external segment of the globus pallidus in primate. J. Comp. Neurol. 2000, 417, 17–31. [Google Scholar] [CrossRef]
  35. Avecillas-Chasin, J.M.; Honey, C.R. Modulation of nigrofugal and pallidofugal pathways in deep brain stimulation for parkinson disease. Neurosurgery 2020, 86, E387–E397. [Google Scholar] [CrossRef] [PubMed]
  36. Saunders, A.; Oldenburg, I.A.; Berezovskii, V.K.; Johnson, C.A.; Kingery, N.D.; Elliott, H.L.; Xie, T.; Gerfen, C.R.; Sabatini, B.L. A direct GABAergic output from the basal ganglia to frontal cortex. Nature 2015, 521, 85–89. [Google Scholar] [CrossRef] [PubMed]
  37. Nambu, A.; Tokuno, H.; Takada, M. Functional significance of the cortico–subthalamo–pallidal ‘hyperdirect’pathway. Neurosci. Res. 2002, 43, 111–117. [Google Scholar] [CrossRef]
  38. Apps, R.; Hawkes, R.; Aoki, S.; Bengtsson, F.; Brown, A.M.; Chen, G.; Ebner, T.J.; Isope, P.; Jörntell, H.; Lackey, E.P. Cerebellar modules and their role as operational cerebellar processing units. Cerebellum 2018, 17, 654–682. [Google Scholar] [CrossRef] [Green Version]
  39. Narain, D.; Remington, E.D.; De Zeeuw, C.I.; Jazayeri, M. A cerebellar mechanism for learning prior distributions of time intervals. Nat. Commun. 2018, 9, 469. [Google Scholar] [CrossRef] [Green Version]
  40. Dum, R.P.; Strick, P.L. An unfolded map of the cerebellar dentate nucleus and its projections to the cerebral cortex. J. Neurophysiol. 2003, 89, 634–639. [Google Scholar] [CrossRef]
  41. Bostan, A.C.; Strick, P.L. The basal ganglia and the cerebellum: Nodes in an integrated network. Nat. Rev. Neurosci. 2018, 19, 338–350. [Google Scholar] [CrossRef]
  42. Lehéricy, S.; Benali, H.; Van de Moortele, P.-F.; Pélégrini-Issac, M.; Waechter, T.; Ugurbil, K.; Doyon, J. Distinct basal ganglia territories are engaged in early and advanced motor sequence learning. Proc. Natl. Acad. Sci. USA 2005, 102, 12566–12571. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Hoshi, E.; Tremblay, L.; Féger, J.; Carras, P.L.; Strick, P.L. The cerebellum communicates with the basal ganglia. Nat. Neurosci. 2005, 8, 1491–1493. [Google Scholar] [CrossRef] [PubMed]
  44. Buisseret-Delmas, C.; Angaut, P. The cerebellar olivo-corticonuclear connections in the rat. Prog. Neurobiol. 1993, 40, 63–87. [Google Scholar] [CrossRef]
  45. Hilker, R.; Voges, J.; Weisenbach, S.; Kalbe, E.; Burghaus, L.; Ghaemi, M.; Lehrke, R.; Koulousakis, A.; Herholz, K.; Sturm, V. Subthalamic nucleus stimulation restores glucose metabolism in associative and limbic cortices and in cerebellum: Evidence from a FDG-PET study in advanced Parkinson’s disease. J. Cereb. Blood Flow Metab. 2004, 24, 7–16. [Google Scholar] [CrossRef] [Green Version]
  46. Castrioto, A.; Moro, E. New targets for deep brain stimulation treatment of Parkinson’s disease. Expert Rev. Neurother. 2013, 13, 1319–1328. [Google Scholar] [CrossRef]
  47. Pahapill, P.A.; Lozano, A.M. The pedunculopontine nucleus and Parkinson’s disease. Brain 2000, 123, 1767–1783. [Google Scholar] [CrossRef]
  48. Alam, M.; Schwabe, K.; Krauss, J.K. The pedunculopontine nucleus area: Critical evaluation of interspecies differences relevant for its use as a target for deep brain stimulation. Brain 2011, 134, 11–23. [Google Scholar] [CrossRef] [Green Version]
  49. Nandi, D.; Stein, J.; Aziz, T. Exploration of the role of the upper brainstem in motor control. Stereotact. Funct. Neurosurg. 2002, 78, 158–167. [Google Scholar] [CrossRef]
  50. Jenkinson, N.; Nandi, D.; Miall, R.C.; Stein, J.F.; Aziz, T.Z. Pedunculopontine nucleus stimulation improves akinesia in a Parkinsonian monkey. Neuroreport 2004, 15, 2621–2624. [Google Scholar] [CrossRef] [Green Version]
  51. Goetz, L.; Bhattacharjee, M.; Ferraye, M.U.; Fraix, V.; Maineri, C.; Nosko, D.; Fenoy, A.J.; Piallat, B.; Torres, N.; Krainik, A. Deep brain stimulation of the pedunculopontine nucleus area in Parkinson disease: MRI-based anatomoclinical correlations and optimal target. Neurosurgery 2019, 84, 506–518. [Google Scholar] [CrossRef] [Green Version]
  52. Ferraye, M.U.; Debû, B.; Fraix, V.; Goetz, L.; Ardouin, C.; Yelnik, J.; Henry-Lagrange, C.; Seigneuret, E.; Piallat, B.; Krack, P.; et al. Effects of pedunculopontine nucleus area stimulation on gait disorders in Parkinson’s disease. Brain 2010, 133, 205–214. [Google Scholar] [CrossRef] [PubMed]
  53. Tykocki, T.; Mandat, T.; Nauman, P. Pedunculopontine nucleus deep brain stimulation in Parkinson’s disease. Arch. Med. Sci. AMS 2011, 7, 555–564. [Google Scholar] [CrossRef]
  54. Stefani, A.; Lozano, A.M.; Peppe, A.; Stanzione, P.; Galati, S.; Tropepi, D.; Pierantozzi, M.; Brusa, L.; Scarnati, E.; Mazzone, P. Bilateral deep brain stimulation of the pedunculopontine and subthalamic nuclei in severe Parkinson’s disease. Brain 2007, 130, 1596–1607. [Google Scholar] [CrossRef] [PubMed]
  55. Plaha, P.; Javed, S.; Agombar, D.; O’Farrell, G.; Khan, S.; Whone, A.; Gill, S. Bilateral caudal zona incerta nucleus stimulation for essential tremor: Outcome and quality of life. J. Neurol. Neurosurg. Psychiatry 2011, 82, 899–904. [Google Scholar] [CrossRef] [PubMed]
  56. Godinho, F.; Thobois, S.; Magnin, M.; Guenot, M.; Polo, G.; Benatru, I.; Xie, J.; Salvetti, A.; Garcia-Larrea, L.; Broussolle, E.; et al. Subthalamic nucleus stimulation in Parkinson’s disease. J. Neurol. 2006, 253, 1347–1355. [Google Scholar] [CrossRef] [PubMed]
  57. Miocinovic, S.; Parent, M.; Butson, C.R.; Hahn, P.J.; Russo, G.S.; Vitek, J.L.; McIntyre, C.C. Computational analysis of subthalamic nucleus and lenticular fasciculus activation during therapeutic deep brain stimulation. J. Neurophysiol. 2006, 96, 1569–1580. [Google Scholar] [CrossRef] [Green Version]
  58. Butson, C.R.; Cooper, S.E.; Henderson, J.M.; McIntyre, C.C. Patient-specific analysis of the volume of tissue activated during deep brain stimulation. Neuroimage 2007, 34, 661–670. [Google Scholar] [CrossRef] [Green Version]
  59. Hägglund, P.; Sandström, L.; Blomstedt, P.; Karlsson, F. Voice Tremor in Patients with Essential Tremor: Effects of Deep Brain Stimulation of Caudal Zona Incerta. J. Voice 2016, 30, 228–233. [Google Scholar] [CrossRef]
  60. Wang, X.; Chou, X.L.; Zhang, L.I.; Tao, H.W. Zona Incerta: An Integrative Node for Global Behavioral Modulation. Trends Neurosci. 2020, 43, 82–87. [Google Scholar] [CrossRef]
  61. dos Santos Ghilardi, M.G.; Cury, R.G.; dos Ângelos, J.S.; Barbosa, D.C.; Barbosa, E.R.; Teixeira, M.J.; Fonoff, E.T. Long-term improvement of tremor and ataxia after bilateral DBS of VoP/zona incerta in FXTAS. Neurology 2015, 84, 1904–1906. [Google Scholar] [CrossRef]
  62. Ossowska, K. Zona incerta as a therapeutic target in Parkinson’s disease. J. Neurol. 2020, 267, 591–606. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Eisinger, R.S.; Wong, J.; Almeida, L.; Ramirez-Zamora, A.; Cagle, J.N.; Giugni, J.C.; Ahmed, B.; Bona, A.R.; Monari, E.; Wagle Shukla, A.; et al. Ventral Intermediate Nucleus Versus Zona Incerta Region Deep Brain Stimulation in Essential Tremor. Mov. Disord. Clin. Pract. 2017, 5, 75–82. [Google Scholar] [CrossRef] [PubMed]
  64. Sasagawa, A.; Enatsu, R.; Kitagawa, M.; Mikami, T.; Nakayama-Kamada, C.; Kuribara, T.; Hirano, T.; Arihara, M.; Mikuni, N. Target Selection of Directional Lead in Patients with Parkinson’s Disease. Neurol. Med. Chir. 2020, 60, 622–628. [Google Scholar] [CrossRef]
  65. Benabid, A.L.; Pollak, P.; Hoffmann, D.; Gervason, C.; Hommel, M.; Perret, J.E.; De Rougemont, J.; Gao, D.M. Long-term suppression of tremor by chronic stimulation of the ventral intermediate thalamic nucleus. Lancet. 1991, 337, 403–406. [Google Scholar] [CrossRef]
  66. Warren, A.E.; Dalic, L.J.; Thevathasan, W.; Roten, A.; Bulluss, K.J.; Archer, J. Targeting the centromedian thalamic nucleus for deep brain stimulation. J. Neurol. Neurosurg. Psychiatry 2020, 91, 339–349. [Google Scholar] [CrossRef] [PubMed]
  67. Kerkerian-Le Goff, L.; Jouve, L.; Melon, C.; Salin, P. Rationale for targeting the thalamic centre-median parafascicular complex in the surgical treatment of Parkinson’s disease. Parkinsonism Relat. Disord. 2009, 15, S167–S170. [Google Scholar] [CrossRef]
  68. Caparros-Lefebvre, D.; Blond, S.; Feltin, M.-P.; Pollak, P.; Benabid, A.L. Improvement of levodopa induced dyskinesias by thalamic deep brain stimulation is related to slight variation in electrode placement: Possible involvement of the centre median and parafascicularis complex. J. Neurol. Neurosurg. Psychiatry 1999, 67, 308–314. [Google Scholar] [CrossRef]
  69. Peppe, A.; Gasbarra, A.; Stefani, A.; Chiavalon, C.; Pierantozzi, M.; Fermi, E.; Stanzione, P.; Caltagirone, C.; Mazzone, P. Deep brain stimulation of CM/PF of thalamus could be the new elective target for tremor in advanced Parkinson’s Disease? Parkinsonism Relat. Disord. 2008, 14, 501–504. [Google Scholar] [CrossRef]
  70. Stefani, A.; Peppe, A.; Pierantozzi, M.; Galati, S.; Moschella, V.; Stanzione, P.; Mazzone, P. Multi-target strategy for Parkinsonian patients: The role of deep brain stimulation in the centromedian–parafascicularis complex. Brain Res. Bull. 2009, 78, 113–118. [Google Scholar] [CrossRef]
  71. Jouve, L.; Salin, P.; Melon, C.; Kerkerian-Le Goff, L. Deep brain stimulation of the center median–parafascicular complex of the thalamus has efficient anti-parkinsonian action associated with widespread cellular responses in the basal ganglia network in a rat model of Parkinson’s disease. J. Neurosci. 2010, 30, 9919–9928. [Google Scholar] [CrossRef] [Green Version]
  72. Takakusaki, K.; Habaguchi, T.; Ohtinata-Sugimoto, J.; Saitoh, K.; Sakamoto, T. Basal ganglia efferents to the brainstem centers controlling postural muscle tone and locomotion: A new concept for understanding motor disorders in basal ganglia dysfunction. Neuroscience 2003, 119, 293–308. [Google Scholar] [CrossRef] [Green Version]
  73. Anderson, D.; Beecher, G.; Ba, F. Deep brain stimulation in Parkinson’s disease: New and emerging targets for refractory motor and nonmotor symptoms. Parkinsons Dis. 2017, 5, 5124328. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Chastan, N.; Westby, G.W.M.; Yelnik, J.; Bardinet, E.; Do, M.C.; Agid, Y.; Welter, M.L. Effects of nigral stimulation on locomotion and postural stability in patients with Parkinson’s disease. Brain 2009, 132, 172–184. [Google Scholar] [CrossRef] [PubMed]
  75. Weiss, D.; Walach, M.; Meisner, C.; Fritz, M.; Scholten, M.; Breit, S.; Plewnia, C.; Bender, B.; Gharabaghi, A.; Wächter, T.; et al. Nigral stimulation for resistant axial motor impairment in Parkinson’s disease? A randomized controlled trial. Brain 2013, 136, 2098–2108. [Google Scholar] [CrossRef] [Green Version]
  76. Valldeoriola, F.; Muñoz, E.; Rumià, J.; Roldán, P.; Cámara, A.; Compta, Y.; Martí, M.J.; Tolosa, E. Simultaneous low-frequency deep brain stimulation of the substantia nigra pars reticulata and high-frequency stimulation of the subthalamic nucleus to treat levodopa unresponsive freezing of gait in Parkinson’s disease: A pilot study. Parkinsonism Relat. Disord. 2019, 60, 153–157. [Google Scholar] [CrossRef] [Green Version]
  77. Parker, T.; Raghu, A.L.; FitzGerald, J.J.; Green, A.L.; Aziz, T.Z. Multitarget deep brain stimulation for clinically complex movement disorders. J. Neurosurg. 2020, 1, 351–356. [Google Scholar] [CrossRef]
  78. Slotty, P.J.; Poologaindran, A.; Honey, C.R. A prospective, randomized, blinded assessment of multitarget thalamic and pallidal deep brain stimulation in a case of hemidystonia. Clin. Neurol. Neurosurg. 2015, 138, 16–19. [Google Scholar] [CrossRef]
  79. Mazzone, P. Deep brain stimulation in Parkinson’s disease: Bilateral implantation of globus pallidus and subthalamic nucleus. J. Neurosurg. Sci. 2003, 47, 47–51. [Google Scholar]
  80. Mazzone, P.; Brown, P.; DiLazzaro, V.; Stanzione, P.; Oliviero, A.; Peppe, A.; Santilli, V.; Insola, A.; Altibrandi, M. Bilateral implantation in globus pallidus internus and in subthalamic nucleus in Parkinson’s disease. Neuromodul. Technol. Neural Interface 2005, 8, 1–6. [Google Scholar] [CrossRef]
  81. Follett, K.A.; Weaver, F.M.; Stern, M.; Hur, K.; Harris, C.L.; Luo, P.; Marks, W.J., Jr.; Rothlind, J.; Sagher, O.; Moy, C.; et al. Pallidal versus subthalamic deep-brain stimulation for Parkinson’s disease. N. Engl. J. Med. 2010, 362, 2077–2091. [Google Scholar] [CrossRef] [Green Version]
  82. Oxley, T.J.; Opie, N.L.; John, S.E.; Rind, G.S.; Ronayne, S.M.; Wheeler, T.L.; Judy, J.W.; McDonald, A.J.; Dornom, A.; Lovell, T.J.; et al. Minimally invasive endovascular stent-electrode array for high-fidelity, chronic recordings of cortical neural activity. Nat. Biotechnol. 2016, 34, 320–327. [Google Scholar] [CrossRef] [PubMed]
  83. Rajah, G.; Saber, H.; Singh, R.; Rangel-Castilla, L. Endovascular delivery of leads and stentrodes and their applications to deep brain stimulation and neuromodulation: A review. Neurosurg. Focus. 2018, 45, E19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Oxley, T.J.; Yoo, P.E.; Rind, G.S.; Ronayne, S.M.; Lee, C.S.; Bird, C.; Hampshire, V.; Sharma, R.P.; Morokoff, A.; Williams, D.L.; et al. Motor neuroprosthesis implanted with neurointerventional surgery improves capacity for activities of daily living tasks in severe paralysis: First in-human experience. J. Neurointerv. Surg. 2021, 13, 102–108. [Google Scholar] [CrossRef]
  85. Turk, A.; Niemann, D.; Ahmed, A.; Aagaard-Kienitz, B. Use of self-expanding stents in distal small cerebral vessels. Am. J. Neuroradiol. 2007, 28, 533–536. [Google Scholar]
  86. Bernier, M.; Cunnane, S.C.; Whittingstall, K. The morphology of the human cerebrovascular system. Hum. Brain Mapp. 2018, 39, 4962–4975. [Google Scholar] [CrossRef] [Green Version]
  87. Neudorfer, C.; Bhatia, K.; Boutet, A.; Germann, J.; Elias, G.J.; Loh, A.; Paff, M.; Krings, T.; Lozano, A.M. Endovascular deep brain stimulation: Investigating the relationship between vascular structures and deep brain stimulation targets. Brain Stimul. 2020, 13, 1668–1677. [Google Scholar] [CrossRef] [PubMed]
  88. Rangel-Castilla, L.; Cress, M.C.; Munich, S.A.; Sonig, A.; Krishna, C.; Gu, E.Y.; Snyder, K.V.; Hopkins, L.N.; Siddiqui, A.H.; Levy, E.I. Feasibility, safety, and periprocedural complications of pipeline embolization for intracranial aneurysm treatment under conscious sedation: University at Buffalo Neurosurgery Experience. Oper. Neurosurg. 2015, 11, 426–430. [Google Scholar] [CrossRef]
  89. Orlov, K.; Kislitsin, D.; Strelnikov, N.; Berestov, V.; Gorbatykh, A.; Shayakhmetov, T.; Seleznev, P.; Tasenko, A. Experience using pipeline embolization device with Shield Technology in a patient lacking a full postoperative dual antiplatelet therapy regimen. Interv. Neuroradiol. 2018, 24, 270–273. [Google Scholar] [CrossRef]
  90. Jenkinson, N.; Brown, P. New insights into the relationship between dopamine, beta oscillations and motor function. Trends Neurosci. 2011, 34, 611–618. [Google Scholar] [CrossRef]
  91. Hammond, C.; Bergman, H.; Brown, P. Pathological synchronization in Parkinson’s disease: Networks, models and treatments. Trends Neurosci. 2007, 30, 357–364. [Google Scholar] [CrossRef]
  92. Brittain, J.S.; Brown, P. Oscillations and the basal ganglia: Motor control and beyond. Neuroimage 2014, 85, 637–647. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Kühn, A.A.; Kupsch, A.; Schneider, G.H.; Brown, P. Reduction in subthalamic 8–35 Hz oscillatory activity correlates with clinical improvement in Parkinson’s disease. Eur. J. Neurosci. 2006, 23, 1956–1960. [Google Scholar] [CrossRef] [PubMed]
  94. Chen, Y.; Wang, J.; Kang, Y.; Ghori, M.B. Emergence of Beta Oscillations of a Resonance Model for Parkinson’s Disease. Neural Plast. 2020, 2020, 8824760. [Google Scholar] [CrossRef] [PubMed]
  95. Davis, N.J.; van Koningsbruggen, M.G. “Non-invasive” brain stimulation is not non-invasive. Front. Syst. Neurosci. 2013, 7, 76. [Google Scholar] [CrossRef] [PubMed]
  96. Schulz, R.; Gerloff, C.; Hummel, F.C. Non-invasive brain stimulation in neurological diseases. Neuropharmacology 2013, 64, 579–587. [Google Scholar] [CrossRef] [PubMed]
  97. Priori, A.; Hallett, M.; Rothwell, J.C. Repetitive transcranial magnetic stimulation or transcranial direct current stimulation? Brain Stimul. 2009, 2, 241–245. [Google Scholar] [CrossRef] [PubMed]
  98. Liebetanz, D.; Fauser, S.; Michaelis, T.; Czéh, B.; Watanabe, T.; Paulus, W.; Frahm, J.; Fuchs, E. Safety aspects of chronic low-frequency transcranial magnetic stimulation based on localized proton magnetic resonance spectroscopy and histology of the rat brain. J. Psychiatr. Res. 2003, 37, 277–286. [Google Scholar] [CrossRef]
  99. Rezayat, E.; Toostani, I.G. A Review on Brain Stimulation Using Low Intensity Focused Ultrasound. Basic Clin. Neurosci. 2016, 7, 187–194. [Google Scholar]
  100. Montemurro, N.; Perrini, P.; Marani, W.; Chaurasia, B.; Corsalini, M.; Scarano, A.; Rapone, B. Multiple Brain Abscesses of Odontogenic Origin. May Oral Microbiota Affect Their Development? A Review of the Current Literature. Appl. Sci. 2021, 11, 3316. [Google Scholar] [CrossRef]
  101. Bronstein, J.M.; Tagliati, M.; Alterman, R.L.; Lozano, A.M.; Volkmann, J.; Stefani, A.; Horak, F.B.; Okun, M.S.; Foote, K.D.; Krack, P. Deep brain stimulation for Parkinson disease: An expert consensus and review of key issues. Arch. Neurol. 2011, 68, 165. [Google Scholar] [CrossRef]
  102. Das, S.; Montemurro, N.; Ashfaq, M.; Ghosh, D.; Sarker, A.C.; Khan, A.H.; Dey, S.; Chaurasia, B. Resolution of Papilledema Following Ventriculoperitoneal Shunt or Endoscopic Third Ventriculostomy for Obstructive Hydrocephalus: A Pilot Study. Medicina 2022, 58, 281. [Google Scholar] [CrossRef]
  103. Fregni, F.; Simon, D.K.; Wu, A.; Pascual-Leone, A. Non-invasive brain stimulation for Parkinson’s disease: A systematic review and meta-analysis of the literature. J. Neurol. Neurosurg. Psychiatry 2005, 76, 1614–1623. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Benninger, D.H.; Lomarev, M.; Lopez, G.; Wassermann, E.M.; Li, X.; Considine, E.; Hallett, M. Transcranial direct current stimulation for the treatment of Parkinson’s disease. J. Neurol. Neurosurg. Psychiatry 2011, 82, 354. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Latorre, A.; Rocchi, L.; Berardelli, A.; Bhatia, K.P.; Rothwell, J.C. The use of transcranial magnetic stimulation as a treatment for movement disorders: A critical review. Mov. Disord. 2019, 34, 769–782. [Google Scholar] [CrossRef] [PubMed]
  106. Chen, R. Guideline on therapeutic use of repetitive transcranial magnetic stimulation: Useful but know the methods and limitations. Clin. Neurophysiol. 2020, 131, 461–462. [Google Scholar] [CrossRef]
  107. Brandt, S.; Ploner, C.; Meyer, B.U. Repetitive transkranielle Magnetstimulation. Der Nervenarzt 1997, 68, 778–784. [Google Scholar] [CrossRef]
  108. Paulus, W. Transcranial brain stimulation: Potential and limitations. e-Neuroforum 2014, 5, 29–36. [Google Scholar] [CrossRef]
  109. Woods, A.J.; Antal, A.; Bikson, M.; Boggio, P.S.; Brunoni, A.R.; Celnik, P.; Cohen, L.G.; Fregni, F.; Herrmann, C.S.; Kappenman, E.S.; et al. A technical guide to tDCS, and related non-invasive brain stimulation tools. Clin. Neurophysiol. 2016, 127, 1031–1048. [Google Scholar] [CrossRef] [Green Version]
  110. Fritsch, B.; Reis, J.; Martinowich, K.; Schambra, H.M.; Ji, Y.; Cohen, L.G.; Lu, B. Direct current stimulation promotes BDNF-dependent synaptic plasticity: Potential implications for motor learning. Neuron 2010, 66, 198–204. [Google Scholar] [CrossRef] [Green Version]
  111. Brittain, J.S.; Cagnan, H. Recent trends in the use of electrical neuromodulation in Parkinson’s disease. Curr. Behav. Neurosci. Rep. 2018, 5, 170–178. [Google Scholar] [CrossRef] [Green Version]
  112. Krause, V.; Wach, C.; Südmeyer, M.; Ferrea, S.; Schnitzler, A.; Pollok, B. Cortico-muscular coupling and motor performance are modulated by 20 Hz transcranial alternating current stimulation (tACS) in Parkinson’s disease. Front. Hum. Neurosci. 2014, 7, 928. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Brittain, J.S.; Cagnan, H.; Mehta, A.R.; Saifee, T.A.; Edwards, M.J.; Brown, P. Distinguishing the central drive to tremor in Parkinson’s disease and essential tremor. J. Neurosci. 2015, 35, 795–806. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Thut, G.; Bergmann, T.O.; Fröhlich, F.; Soekadar, S.R.; Brittain, J.S.; Valero-Cabré, A.; Sack, A.T.; Miniussi, C.; Antal, A.; Siebner, H.R.; et al. Guiding transcranial brain stimulation by EEG/MEG to interact with ongoing brain activity and associated functions: A position paper. Clin. Neurophysiol. 2017, 128, 843–857. [Google Scholar] [CrossRef] [Green Version]
  115. Del Felice, A.; Castiglia, L.; Formaggio, E.; Cattelan, M.; Scarpa, B.; Manganotti, P.; Tenconi, E.; Masiero, S. Personalized transcranial alternating current stimulation (tACS) and physical therapy to treat motor and cognitive symptoms in Parkinson’s disease: A randomized cross-over trial. NeuroImage Clin. 2019, 22, 101768. [Google Scholar] [CrossRef]
  116. Shill, H.A.; Obradov, S.; Katsnelson, Y.; Pizinger, R. A randomized, double-blind trial of transcranial electrostimulation in early Parkinson’s disease. Mov. Disord. 2011, 26, 1477–1480. [Google Scholar] [CrossRef] [PubMed]
  117. Moosa, S.; Martínez-Fernández, R.; Elias, W.J.; Del Alamo, M.; Eisenberg, H.M.; Fishman, P.S. The role of high-intensity focused ultrasound as a symptomatic treatment for Parkinson’s disease. Mov. Disord. 2019, 34, 1243–1251. [Google Scholar] [CrossRef] [PubMed]
  118. Schlesinger, I.; Sinai, A.; Zaaroor, M. MRI-guided focused ultrasound in Parkinson’s disease: A review. Parkinsons Dis. 2017, 2017, 8124624. [Google Scholar] [CrossRef]
  119. Quadri, S.A.; Waqas, M.; Khan, I.; Khan, M.A.; Suriya, S.S.; Farooqui, M.; Fiani, B. High-intensity focused ultrasound: Past, present, and future in neurosurgery. Neurosurg. Focus 2018, 44, E16. [Google Scholar] [CrossRef] [Green Version]
  120. Zaaroor, M.; Sinai, A.; Goldsher, D.; Eran, A.; Nassar, M.; Schlesinger, I. Magnetic resonance-guided focused ultrasound thalamotomy for tremor: A report of 30 Parkinson’s disease and essential tremor cases. J Neurosurg. 2018, 128, 202–210. [Google Scholar] [CrossRef] [Green Version]
  121. Chung, A.H.; Jolesz, F.A.; Hynynen, K. Thermal dosimetry of a focused ultrasound beam in vivo by magnetic resonance imaging. Med. Phys. 1999, 26, 2017–2026. [Google Scholar] [CrossRef]
  122. Burks, S.R.; Ziadloo, A.; Hancock, H.A.; Chaudhry, A.; Dean, D.D.; Lewis, B.K.; Frenkel, V.; Frank, J.A. Investigation of cellular and molecular responses to pulsed focused ultrasound in a mouse model. PLoS ONE 2011, 6, e24730. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Krishna, V.; Sammartino, F.; Rezai, A. A review of the current therapies, challenges, and future directions of transcranial focused ultrasound technology: Advances in diagnosis and treatment. JAMA Neurol. 2018, 75, 246–254. [Google Scholar] [CrossRef] [PubMed]
  124. LeWitt, P.A.; Lipsman, N.; Kordower, J.H. Focused ultrasound opening of the blood-brain barrier for treatment of Parkinson’s disease. Mov. Disord. 2019, 34, 1274–1278. [Google Scholar] [CrossRef] [PubMed]
  125. Martínez-Fernández, R.; Máñez-Miró, J.U.; Rodríguez-Rojas, R.; Del Álamo, M.; Shah, B.B.; Hernández-Fernández, F.; Pineda-Pardo, J.A.; Monje, M.H.; Fernández-Rodríguez, B.; Sperling, S.A.; et al. Randomized trial of focused ultrasound subthalamotomy for Parkinson’s disease. N. Engl. J. Med. 2020, 383, 2501–2513. [Google Scholar] [CrossRef]
  126. Jung, N.Y.; Park, C.K.; Kim, M.; Lee, P.H.; Sohn, Y.H.; Chang, J.W. The efficacy and limits of magnetic resonance-guided focused ultrasound pallidotomy for Parkinson’s disease: A Phase I clinical trial. J Neurosurg. 2018, 130, 1853–1861. [Google Scholar] [CrossRef] [Green Version]
  127. Tan, E.K.; Chan, L.L. A Trial of focused ultrasound thalamotomy for essential tremor. N. Engl. J. Med. 2016, 375, 2202–2203. [Google Scholar]
  128. Martin, E.; Jeanmonod, D.; Morel, A.; Zadicario, E.; Werner, B. High-intensity focused ultrasound for noninvasive functional neurosurgery. Ann. Neurol. 2009, 66, 858–861. [Google Scholar] [CrossRef] [Green Version]
  129. Bond, A.E.; Shah, B.B.; Huss, D.S.; Dallapiazza, R.F.; Warren, A.; Harrison, M.B.; Sperling, S.A.; Wang, X.Q.; Gwinn, R.; Witt, J.; et al. Safety and Efficacy of focused ultrasound thalamotomy for patients with medication-refractory, tremor-dominant Parkinson disease: A randomized clinical trial. JAMA Neurol. 2017, 74, 1412–1418. [Google Scholar] [CrossRef]
  130. Ohye, C.; Higuchi, Y.; Shibazaki, T.; Hashimoto, T.; Koyama, T.; Hirai, T.; Matsuda, S.; Serizawa, T.; Hori, T.; Hayashi, M.; et al. Gamma knife thalamotomy for Parkinson disease and essential tremor: A prospective multicenter study. Neurosurgery 2012, 70, 526–536. [Google Scholar] [CrossRef]
  131. Dauer, W.; Przedborski, S. Parkinson’s disease: Mechanisms and models. Neuron 2003, 39, 889–909. [Google Scholar] [CrossRef] [Green Version]
  132. Harris, J.P.; Burrell, J.C.; Struzyna, L.A.; Chen, H.I.; Serruya, M.D.; Wolf, J.A.; Duda, J.E.; Cullen, D.K. Emerging regenerative medicine and tissue engineering strategies for Parkinson’s disease. NPJ Parkinsons Dis. 2020, 6, 4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Johnson, S.J.; Diener, M.D.; Kaltenboeck, A.; Birnbaum, H.G.; Siderowf, A.D. An economic model of Parkinson’s disease: Implications for slowing progression in the United States. Mov. Disord. 2013, 28, 319–326. [Google Scholar] [CrossRef] [PubMed]
  134. Khademhosseini, A.; Langer, R. A decade of progress in tissue engineering. Nat. Protoc. 2016, 11, 1775–1781. [Google Scholar] [CrossRef] [PubMed]
  135. Petersen, S.E.; Sporns, O. Brain Networks and Cognitive Architectures. Neuron 2015, 88, 207–219. [Google Scholar] [CrossRef] [Green Version]
  136. Björklund, A.; Schmidt, R.H.; Stenevi, U. Functional reinnervation of the neostriatum in the adult rat by use of intraparenchymal grafting of dissociated cell suspensions from the substantia nigra. Cell Tissue Res. 1980, 212, 39–45. [Google Scholar] [CrossRef]
  137. Isacson, O.; Bjorklund, L.M.; Schumacher, J.M. Toward full restoration of synaptic and terminal function of the dopaminergic system in Parkinson’s disease by stem cells. Ann. Neurol. 2003, 53, S135–S148. [Google Scholar] [CrossRef]
  138. Rivetti Di Val Cervo, P.; Romanov, R.A.; Spigolon, G.; Masini, D.; Martín-Montañez, E.; Toledo, E.M.; La Manno, G.; Feyder, M.; Pifl, C.; Ng, Y.H.; et al. Induction of functional dopamine neurons from human astrocytes in vitro and mouse astrocytes in a Parkinson’s disease model. Nat. Biotechnol. 2017, 35, 444–452. [Google Scholar] [CrossRef]
  139. Harris, J.P.; Struzyna, L.A.; Murphy, P.L.; Adewole, D.O.; Kuo, E.; Cullen, D.K. Advanced biomaterial strategies to transplant preformed micro-tissue engineered neural networks into the brain. J. Neural Eng. 2016, 13, 016019. [Google Scholar] [CrossRef] [Green Version]
  140. Li, H.; Chen, G. In Vivo Reprogramming for CNS Repair: Regenerating Neurons from Endogenous Glial Cells. Neuron 2016, 91, 728–738. [Google Scholar] [CrossRef] [Green Version]
  141. Dunnett, S.B.; Bjorklund, A.; Schmidt, R.H.; Stenevi, U.; Iversen, S.D. Intracerebral grafting of neuronal cell suspensions. IV. Behavioural recovery in rats with unilateral 6-OHDA lesions following implantation of nigral cell suspensions in different forebrain sites. Acta Physiol. Scand 1983, 522, 29–37. [Google Scholar]
  142. Björklund, A.; Lindvall, O. Cell replacement therapies for central nervous system disorders. Nat. Neurosci. 2000, 3, 537–544. [Google Scholar] [CrossRef] [PubMed]
  143. Stoker, T.B.; Torsney, K.M.; Barker, R.A. Emerging Treatment Approaches for Parkinson’s Disease. Front. Neurosci. 2018, 12, 693. [Google Scholar] [CrossRef] [Green Version]
  144. Brundin, P.; Nilsson, O.G.; Strecker, R.E.; Lindvall, O.; Astedt, B.; Björklund, A. Behavioural effects of human fetal dopamine neurons grafted in a rat model of Parkinson’s disease. Exp. Brain Res. 1986, 65, 235–240. [Google Scholar] [CrossRef] [PubMed]
  145. Kordower, J.H.; Chu, Y.; Hauser, R.A.; Freeman, T.B.; Olanow, C.W. Lewy body-like pathology in long-term embryonic nigral transplants in Parkinson’s disease. Nat. Med. 2008, 14, 504–506. [Google Scholar] [CrossRef] [PubMed]
  146. Barker, R.A.; Barrett, J.; Mason, S.L.; Björklund, A. Fetal dopaminergic transplantation trials and the future of neural grafting in Parkinson’s disease. Lancet Neurol. 2013, 12, 84–91. [Google Scholar] [CrossRef]
  147. Bolam, J.P.; Freund, T.F.; Björklund, A.; Dunnett, S.B.; Smith, A.D. Synaptic input and local output of dopaminergic neurons in grafts that functionally reinnervate the host neostriatum. Exp. Brain Res. 1987, 68, 131–146. [Google Scholar] [CrossRef]
  148. Kordower, J.H.; Freeman, T.B.; Chen, E.Y.; Mufson, E.J.; Sanberg, P.R.; Hauser, R.A.; Snow, B.; Warren Olanow, C. Fetal nigral grafts survive and mediate clinical benefit in a patient with Parkinson’s disease. Mov. Disord. 1998, 13, 383–393. [Google Scholar] [CrossRef]
  149. Lindvall, O.; Björklund, A. Cell therapy in Parkinson’s disease. NeuroRx 2004, 1, 382–393. [Google Scholar] [CrossRef]
  150. Olanow, C.W.; Kordower, J.H.; Freeman, T.B. Fetal nigral transplantation as a therapy for Parkinson’s disease. Trends Neurosci. 1996, 19, 102–109. [Google Scholar] [CrossRef]
  151. Li, W.; Englund, E.; Widner, H.; Mattsson, B.; van Westen, D.; Lätt, J.; Rehncrona, S.; Brundin, P.; Björklund, A.; Lindvall, O.; et al. Extensive graft-derived dopaminergic innervation is maintained 24 years after transplantation in the degenerating parkinsonian brain. Proc. Natl. Acad. Sci. USA 2016, 113, 6544–6549. [Google Scholar] [CrossRef] [Green Version]
  152. Brederlau, A.; Correia, A.S.; Anisimov, S.V.; Elmi, M.; Paul, G.; Roybon, L.; Morizane, A.; Bergquist, F.; Riebe, I.; Nannmark, U.; et al. Transplantation of human embryonic stem cell-derived cells to a rat model of Parkinson’s disease: Effect of in vitro differentiation on graft survival and teratoma formation. Stem Cells 2006, 24, 1433–1440. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Lindvall, O.; Björklund, A. Cell therapeutics in Parkinson’s disease. Neurotherapeutics 2011, 8, 539–548. [Google Scholar] [CrossRef] [Green Version]
  154. Bezard, E.; Brotchie, J.M.; Gross, C.E. Pathophysiology of levodopa-induced dyskinesia: Potential for new therapies. Nat. Rev. Neurosci. 2001, 2, 577–588. [Google Scholar] [CrossRef] [PubMed]
  155. Barker, R.A.; Drouin-Ouellet, J.; Parmar, M. Cell-based therapies for Parkinson disease—Past insights and future potential. Nat. Rev. Neurol. 2015, 11, 492–503. [Google Scholar] [CrossRef] [PubMed]
  156. Parmar, M.; Torper, O.; Drouin-Ouellet, J. Cell-based therapy for Parkinson’s disease: A journey through decades toward the light side of the Force. Eur. J. Neurosci. 2019, 49, 463–471. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Grealish, S.; Diguet, E.; Kirkeby, A.; Mattsson, B.; Heuer, A.; Bramoulle, Y.; Van Camp, N.; Perrier, A.L.; Hantraye, P.; Björklund, A.; et al. Human ESC-derived dopamine neurons show similar preclinical efficacy and potency to fetal neurons when grafted in a rat model of Parkinson’s disease. Cell Stem Cell 2014, 15, 653–665. [Google Scholar] [CrossRef] [Green Version]
  158. Yi, B.R.; Kim, S.U.; Choi, K.C. Development and application of neural stem cells for treating various human neurological diseases in animal models. Lab. Anim. Res. 2013, 29, 131–137. [Google Scholar] [CrossRef] [Green Version]
  159. Hellmann, M.A.; Panet, H.; Barhum, Y.; Melamed, E.; Offen, D. Increased survival and migration of engrafted mesenchymal bone marrow stem cells in 6-hydroxydopamine-lesioned rodents. Neurosci. Lett. 2006, 395, 124–128. [Google Scholar] [CrossRef]
  160. Struzyna, L.A.; Katiyar, K.; Cullen, D.K. Living scaffolds for neuroregeneration. Curr. Opin. Solid State Mater Sci. 2014, 18, 308–318. [Google Scholar] [CrossRef] [Green Version]
  161. Montemurro, N.; Scerrati, A.; Ricciardi, L.; Trevisi, G. The Exoscope in Neurosurgery: An Overview of the Current Lit-erature of Intraoperative Use in Brain and Spine Surgery. J. Clin. Med. 2021, 11, 223. [Google Scholar] [CrossRef]
  162. Mishra, R.; Narayanan, M.K.; Umana, G.E.; Montemurro, N.; Chaurasia, B.; Deora, H. Virtual Reality in Neurosurgery: Beyond Neurosurgical Planning. Int. J. Environ. Res. Public Health 2022, 19, 1719. [Google Scholar] [CrossRef] [PubMed]
  163. Montemurro, N. Telemedicine: Could it represent a new problem for spine surgeons to solve? Global Spine J. 2022, 12, 1306–1307. [Google Scholar] [CrossRef] [PubMed]
  164. Winter, C.C.; Katiyar, K.S.; Hernandez, N.S.; Song, Y.J.; Struzyna, L.A.; Harris, J.P.; Cullen, D.K. Transplantable living scaffolds comprised of micro-tissue engineered aligned astrocyte networks to facilitate central nervous system regeneration. Acta Biomater. 2016, 38, 44–58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Ghosh, B.; Zhang, C.; Smith, G.M. Bridging between transplantation therapy and neurotrophic factors in Parkinson’s disease. Front. Biosci. 2014, 6, 225–235. [Google Scholar] [CrossRef]
  166. Ricciardi, L.; Pucci, R.; Piazza, A.; Lofrese, G.; Scerrati, A.; Montemurro, N.; Raco, A.; Miscusi, M.; Ius, T.; Zeppieri, M.; et al. Role of stem cells-based in facial nerve reanimation: A meta-analysis of histological and neurophysiological outcomes. World J. Stem Cells 2022, 14, 420–428. [Google Scholar] [CrossRef]
  167. Montemurro, N.; Condino, S.; Carbone, M.; Cattari, N.; D’Amato, R.; Cutolo, F.; Ferrari, V. Brain Tumor and Augmented Reality: New Technologies for the Future. Int. J. Environ. Res. Public Health 2022, 19, 6347. [Google Scholar] [CrossRef]
  168. Gaillard, A.; Jaber, M. Rewiring the brain with cell transplantation in Parkinson’s disease. Trends Neurosci. 2011, 34, 124–133. [Google Scholar] [CrossRef]
  169. Weng, S.J.; Li, I.H.; Huang, Y.S.; Chueh, S.H.; Chou, T.K.; Huang, S.Y.; Shiue, C.Y.; Cheng, C.Y.; Ma, K.H. KA-bridged transplantation of mesencephalic tissue and olfactory ensheathing cells in a Parkinsonian rat model. J. Tissue Eng. Regen. Med. 2017, 11, 2024–2033. [Google Scholar] [CrossRef]
  170. Cullen, D.K.; Tang-Schomer, M.D.; Struzyna, L.A.; Patel, A.R.; Johnson, V.E.; Wolf, J.A.; Smith, D.H. Microtissue engineered constructs with living axons for targeted nervous system reconstruction. Tissue Eng. Part A 2012, 18, 2280–2289. [Google Scholar] [CrossRef] [Green Version]
  171. Ward, R.J.; Zucca, F.A.; Duyn, J.H.; Crichton, R.R.; Zecca, L. The role of iron in brain ageing and neurodegenerative disorders. Lancet Neurol. 2014, 13, 1045–1060. [Google Scholar] [CrossRef] [Green Version]
  172. Zhang, Q.; Li, C.; Zhang, T.; Ge, Y.; Han, X.; Sun, S.; Ding, J.; Lu, M.; Hu, G. Deletion of Kir6.2/SUR1 potassium channels rescues diminishing of DA neurons via decreasing iron accumulation in PD. Mol. Cell Neurosci. 2018, 92, 164–176. [Google Scholar] [CrossRef] [PubMed]
  173. Yu, C.; Cassar, I.R.; Sambangi, J.; Grill, W.M. Frequency-specific optogenetic deep brain stimulation of subthalamic nucleus improves Parkinsonian motor behaviors. J. Neurosci. 2020, 40, 4323–4334. [Google Scholar] [CrossRef] [PubMed]
  174. Gittis, A.H.; Yttri, E.A. Translating Insights from Optogenetics to Therapies for Parkinson’s Disease. Curr. Opin. Biomed Eng. 2018, 8, 14–19. [Google Scholar] [CrossRef]
  175. Deisseroth, K.; Feng, G.; Majewska, A.K.; Miesenböck, G.; Ting, A.; Schnitzer, M.J. Next-generation optical technologies for illuminating genetically targeted brain circuits. J. Neurosci. 2006, 26, 10380–10386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Gradinaru, V.; Mogri, M.; Thompson, K.R.; Henderson, J.M.; Deisseroth, K. Optical deconstruction of parkinsonian neural circuitry. Science 2009, 324, 354–359. [Google Scholar] [CrossRef] [Green Version]
  177. Rossi, M.A.; Calakos, N.; Yin, H.H. Spotlight on movement disorders: What optogenetics has to offer. Mov. Disord. 2015, 30, 624–631. [Google Scholar] [CrossRef] [Green Version]
  178. Vazey, E.M.; Aston-Jones, G. New tricks for old dogmas: Optogenetic and designer receptor insights for Parkinson’s disease. Brain Res. 2013, 1511, 153–163. [Google Scholar] [CrossRef] [Green Version]
  179. Herman, A.M.; Huang, L.; Murphey, D.K.; Garcia, I.; Arenkiel, B.R. Cell type-specific and time-dependent light exposure contribute to silencing in neurons expressing Channelrhodopsin-2. Elife 2014, 3, e01481. [Google Scholar] [CrossRef]
  180. Raimondo, J.V.; Kay, L.; Ellender, T.J.; Akerman, C.J. Optogenetic silencing strategies differ in their effects on inhibitory synaptic transmission. Nat. Neurosci. 2012, 15, 1102–1104. [Google Scholar] [CrossRef] [Green Version]
  181. Edgerton, J.R.; Jaeger, D. Optogenetic activation of nigral inhibitory inputs to motor thalamus in the mouse reveals classic inhibition with little potential for rebound activation. Front. Cell. Neurosci. 2014, 8, 36. [Google Scholar] [CrossRef] [Green Version]
Figure 1. This figure shows a coronal section of the basal ganglia circuit with emphasis on its complex connections. Illustrated by J. Lizana.
Figure 1. This figure shows a coronal section of the basal ganglia circuit with emphasis on its complex connections. Illustrated by J. Lizana.
Ijerph 19 08799 g001
Figure 2. Figure (a) depicts the classical pallidofugal pathways, while Figure (b) illustrates some changes in the recent understanding of how the Globus pallidus communicates with other basal ganglia structures. Illustrated by J. Lizana.
Figure 2. Figure (a) depicts the classical pallidofugal pathways, while Figure (b) illustrates some changes in the recent understanding of how the Globus pallidus communicates with other basal ganglia structures. Illustrated by J. Lizana.
Ijerph 19 08799 g002
Figure 3. A sagittal section of the brain, trunk and cerebellum shows the complex connections between basal ganglia and cerebellum. AMY, Amygdala; BC, Brain cortex; CC, Cerebellar cortex; CiC, Cingulate cortex; CN, Caudate nucleus; DN, Dentate nucleus; FN, Fastigial nucleus; GPe, External globus pallidus; GPi, Internal globus pallidus; IN, Interpositus nucleus; NAcc, Nucleus accumbens; NBM, Nucleus basalis of Meynert; PN, Pontine nuclei; PUT, Putamen; RN, Red nucleus; STH, Subthalamus; THA, Thalamus. Illustrated by J. Lizana.
Figure 3. A sagittal section of the brain, trunk and cerebellum shows the complex connections between basal ganglia and cerebellum. AMY, Amygdala; BC, Brain cortex; CC, Cerebellar cortex; CiC, Cingulate cortex; CN, Caudate nucleus; DN, Dentate nucleus; FN, Fastigial nucleus; GPe, External globus pallidus; GPi, Internal globus pallidus; IN, Interpositus nucleus; NAcc, Nucleus accumbens; NBM, Nucleus basalis of Meynert; PN, Pontine nuclei; PUT, Putamen; RN, Red nucleus; STH, Subthalamus; THA, Thalamus. Illustrated by J. Lizana.
Ijerph 19 08799 g003
Figure 4. This figure shows some theoretical locations (superior sagittal sinus, internal cerebral vein and basal vein of Rosenthal) of STENTrodes which can be placed by the endovascular transvenous approach to stimulate the cortex, thalamus or brainstem nuclei. Illustrated by J. Lizana.
Figure 4. This figure shows some theoretical locations (superior sagittal sinus, internal cerebral vein and basal vein of Rosenthal) of STENTrodes which can be placed by the endovascular transvenous approach to stimulate the cortex, thalamus or brainstem nuclei. Illustrated by J. Lizana.
Ijerph 19 08799 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Montemurro, N.; Aliaga, N.; Graff, P.; Escribano, A.; Lizana, J. New Targets and New Technologies in the Treatment of Parkinson’s Disease: A Narrative Review. Int. J. Environ. Res. Public Health 2022, 19, 8799. https://doi.org/10.3390/ijerph19148799

AMA Style

Montemurro N, Aliaga N, Graff P, Escribano A, Lizana J. New Targets and New Technologies in the Treatment of Parkinson’s Disease: A Narrative Review. International Journal of Environmental Research and Public Health. 2022; 19(14):8799. https://doi.org/10.3390/ijerph19148799

Chicago/Turabian Style

Montemurro, Nicola, Nelida Aliaga, Pablo Graff, Amanda Escribano, and Jafeth Lizana. 2022. "New Targets and New Technologies in the Treatment of Parkinson’s Disease: A Narrative Review" International Journal of Environmental Research and Public Health 19, no. 14: 8799. https://doi.org/10.3390/ijerph19148799

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop