Next Article in Journal
Effect of Pd2Spermine on Mice Brain-Liver Axis Metabolism Assessed by NMR Metabolomics
Next Article in Special Issue
Impact of JH Signaling on Reproductive Physiology of the Classical Insect Model, Rhodnius prolixus
Previous Article in Journal
Fascaplysin Induces Apoptosis and Ferroptosis, and Enhances Anti-PD-1 Immunotherapy in Non-Small Cell Lung Cancer (NSCLC) by Promoting PD-L1 Expression
Previous Article in Special Issue
The Origin, Function, Distribution, Quantification, and Research Advances of Extracellular DNA
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Neuroprotection and Non-Invasive Brain Stimulation: Facts or Fiction?

1
“Aldo Ravelli” Center for Neurotechnology and Experimental Brain Therapeutics, Department of Health Sciences, University of Milan, Via Antonio di Rudinì 8, 20142 Milan, Italy
2
Department of Electronics, Information and Bioengineering, Politecnico di Milano, Piazza Leonardo da Vinci, 32, 20133 Milan, Italy
3
Clinical Neurology Unit, “Azienda Socio-Sanitaria Territoriale Santi Paolo E Carlo”, Department of Health Sciences, University of Milan, Via Antonio di Rudinì 8, 20142 Milan, Italy
4
School of Medicine and Surgery, Milan Center for Neuroscience (NeuroMI), University of Milano-Bicocca, Monza, 20126 Milan, Italy
5
Biochemistry and Molecular Biology Laboratory, Department of Health Sciences, University of Milan, 20142 Milan, Italy
6
Department of Neurology, ASST-Monza, San Gerardo Hospital, Monza, 20126 Milan, Italy
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(22), 13775; https://doi.org/10.3390/ijms232213775
Submission received: 4 October 2022 / Revised: 2 November 2022 / Accepted: 5 November 2022 / Published: 9 November 2022
(This article belongs to the Collection Feature Papers in “Molecular Biology”)

Abstract

:
Non-Invasive Brain Stimulation (NIBS) techniques, such as transcranial Direct Current Stimulation (tDCS) and repetitive Magnetic Transcranial Stimulation (rTMS), are well-known non-pharmacological approaches to improve both motor and non-motor symptoms in patients with neurodegenerative disorders. Their use is of particular interest especially for the treatment of cognitive impairment in Alzheimer’s Disease (AD), as well as axial disturbances in Parkinson’s (PD), where conventional pharmacological therapies show very mild and short-lasting effects. However, their ability to interfere with disease progression over time is not well understood; recent evidence suggests that NIBS may have a neuroprotective effect, thus slowing disease progression and modulating the aggregation state of pathological proteins. In this narrative review, we gather current knowledge about neuroprotection and NIBS in neurodegenerative diseases (i.e., PD and AD), just mentioning the few results related to stroke. As further matter of debate, we discuss similarities and differences with Deep Brain Stimulation (DBS)—induced neuroprotective effects, and highlight possible future directions for ongoing clinical studies.

1. Introduction

Non-Invasive Brain Stimulation (NIBS) techniques, including transcranial Direct Current Stimulation (tDCS), transcranial Alternating Current Stimulation (tACS) and repetitive Transcranial Magnetic Stimulation (rTMS), have been proposed for years to improve both motor and non-motor symptoms in a number of neurological conditions, comprising neurodegenerative disorders as Alzheimer’s (AD) and Parkinson’s Disease (PD) [1,2,3,4,5]. They are safe and promising tools for the modulation of cortical and, probably, sub-cortical activities [6]. A growing body of literature strengthens their use for the treatment of both speech disturbances and axial symptoms of PD (bradykinesia, falls and dysphagia), where conventional pharmacological approaches did not provide long-lasting changes over time [7]. Moreover, they proved a significant effect for the treatment of the so-called “freezing of gait” (FOG), which still remains a challenge for clinicians and neuroscientists [8,9,10,11]. However, there is a substantial lack of papers discussing their putative role as disease-modifying, neuroprotective therapies; this is of key importance because pharmacological treatments show merely a “symptomatic” effect, without any significant interference with disease progression over time [12]. In this review, we encompass current knowledge about NIBS and neuroprotection, discussing novel data and old concepts, both in animal and human models, and highlighting the possible use of these techniques in early phases of the neurodegenerative process. Moreover, we suggest a new view of tDCS and rTMS mechanisms of action, not based on either polarity or frequency-dependence of their after-effects, but on their ability to interfere with pathological protein accumulation and degradation in experimental models of neurodegenerative disorders. As a matter of debate and to give a more complete picture about neurostimulation and neuroprotection, we briefly provide a comparison between NIBS and invasive brain stimulation, as DBS (“Deep Brain Stimulation”), towards the goal of neuroprotection, both in degenerative and non-degenerative disorders; accordingly, DBS has recently demonstrated interesting results in animal models of schizophrenia and depression [13,14].

2. NIBS and Neuroprotection in Parkinson’s Disease

2.1. tDCS

Parkinson’s disease (PD) is the second most common neurodegenerative disorder affecting about 1% of the population >60 years of age [15]. The pathological hallmark of PD is the progressive degeneration of dopaminergic neurons in the substantia nigra and striatum [16], ultimately leading to motor (e.g., tremor, akinesia, rigidity, gait impairments) and non-motor (e.g., anxiety, depression, cognitive deficits) symptoms. However, the underpinning mechanisms remain unclear [17]. Multiple factors may contribute to neuronal damage, both in mutation related and in idiopathic forms of the disease [18], including biochemical factors, causing cellular stress accumulation, due to inflammation, oxidative stress and excitotoxicity, and leading to mitochondrial dysfunction, energy production loss and cell demise (e.g., mitochondrial dysfunction, defective protein degradation, neuroinflammation, oxidative stress, excitotoxicity), all of which are tightly linked to each other [19,20,21]. Available treatments are only symptomatic, and pharmacological therapies lead to several and disabling side effects over time [12].
Several neuromodulation techniques have been suggested, such as complementary treatment approach, both invasive [22,23,24] and non-invasive [25,26]. Among these, transcranial direct current stimulation (tDCS) showed a convincing therapeutic potential, with benefits both on motor and cognitive performances [27,28,29,30,31,32,33]. However, the mechanisms by which tDCS exerts its effects in PD patients are not fully understood, particularly at cellular and molecular levels [34]. Current knowledge suggests that tDCS increases the release of dopamine [35,36,37], modulates alpha-synuclein aggregation and autophagic degradation [34], alters neurotransmitters concentration (e.g., GABA, serotonin, glutamate) [38] and induces anti-apoptotic and anti-inflammatory effects [20,39]. However, these cellular effects have been collected either in vitro or in animal models, but not yet confirmed in human studies. In this scenario, tDCS is likely to enhance the expression of brain-derived neurotrophic factor (BDNF) [40], a neurotransmitter modulator and neurotrophic factor that supports neurogenesis [41,42] and survival of neurons [43]. Therefore, these findings suggest a possible neuroprotective effect of tDCS, which appears to be partially effective in restoring some of the biochemical defects associated with neurodegenerative diseases, as confirmed by animal studies [17,19,44,45] (see Figure 1). Indeed, current knowledge comes from neurotoxin-treated animal models, and shows preliminary and promising results in terms of tDCS-induced antioxidant function and survival of dopaminergic cells from neurotoxin-induced cell death [17,19,44,45].
In PD patients, the alteration of tyrosine hydroxylase activity (TH—an enzyme catalysing the precursor of dopamine, L-DOPA, in dopamine) reduces dopamine (DA) levels [46]. Besides, oxidative stress is increased and antioxidative processes are inhibited [19]. tDCS may have a role in neuronal protection acting on the response against oxidative stress, as suggested by Lee et al., 2019 [17] and Li et al., 2015 [19] (see Table 1). Anodal tDCS applied on mice preserves dopaminergic neurons after the injection of the neurotoxin 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) [17] and increases both DA and TH content [19]. Also, it reduces the decrease of antioxidant enzymes activities (superoxide dismutase, SOD; and glutathione peroxidase, GSH-Px) induced by MPTP, ultimately improving the survival response in the nigral-striatal area. However, antioxidative results might not be a direct effect of such a response and may not be driven directly by the stimulation, but rather a consequence of enhanced secretion of BDNF induced by tDCS, as shown by previous studies [38].
The major target of oxidative stress is the mitochondrion-increased production (or decreased capacity to eliminate) of free radicals, known to induce neuronal death via disruption of mitochondrial function [47]. Several findings suggest a pivotal role of mitochondrial dysfunctions in PD pathogenesis [48]. Lee et al., 2019 [17] demonstrated in mice that anodal tDCS exerts a neuroprotective effect against MPTP toxicity, by normalizing mitophagy activation, enhancing mitochondrial biogenesis and restoring mitochondrial damage (see Table 1) [17,49]. tDCS also decreases the effects of MPTP, i.e., increased expression of mitophagy-related proteins (PTEN-induced putative kinase 1, PINK1; Parkin; and microtubule-associated protein light chain 3, LC3), PINK1/Parkin upregulation and enhanced autophagic flux [17]. Mitochondrial biogenesis-related proteins (peroxisome proliferator-activated receptor γ coactivator, PGC1α; and nuclear respirator factor 1, NRF1) and mitochondrial transcription factor A (TFAM) were increased by tDCS, suggesting that its biogenetic effect might be exerted at the level of transcription and replication of mitochondrial DNA [17]. Also, tDCS recovered an MTPT-induced increase of dynamin-related protein 1 (Drp1) expression, which reflects mitochondrial fragmentation and the release of pro-apoptotic proteins [50].
In PD, oxidative stress (OxS), mitochondrial dysfunction, excitotoxicity, and neuroinflammation are strictly linked to autophagy pathways. Autophagy is a cellular homeostatic process involved in both unspecific bulk degradation (macroautophagy, referred to simply as “autophagy”) of cytosolic proteins, aggregates and organelles [51] but also in specific catabolism (chaperone-mediated autophagy, CMA) of neuropathological proteins, including alpha-synuclein [52]. Defects in macroautophagy and CMA have been shown to play an important role in the pathogenesis of the disease [53]. To date, while no data are available in the literature on the specific effect of electrical stimulation on CMA, with the only exception of a recent study from our group [34], studies have investigated the effect on macroautophagy. Lee et al., 2018 [45] demonstrated that anodal tDCS over the left motor cortical area stabilizes the autophagy processes activated by MPTP-induced toxicity, as showed by microtubule-associated protein 1 light chain 3 (LC3) and AMP-activated protein kinase (AMPK) upregulation, and the mechanistic target of rapamycin (mTOR) and sequestosome1/p62 (p62) downregulation in experimental mice (see Table 1). However, such an effect has not been described for MTPT treatment-free, suggesting that the effect of anodal tDCS might occur under stress conditions. Also, anodal tDCS modulates the MPTP-induced upregulation of α-synuclein in substantia nigra pars compacta, which has been identified as a distinctive marker for PD [54]. As for the antioxidative effects, however, it is still under debate whether these effects on autophagy are a direct effect of tDCS, or rather a result of an increased release of BDNF [55].
Overall, these results represent a theoretical basis for the study of tDCS (anodal polarity) as a potential neuroprotective rather than a symptomatic therapy, as has mostly been considered so far. This application would be of great interest, since there is an absolute unmet need for treatments aiming to halt or restore the disease.

2.2. rTMS

A recent study evidenced a neuroprotective effect of early rTMS, as suggested by its ability to preserve tyrosine hydroxylase- (TH-) positive neurons in the substantia nigra pars compacta (SNpc) and fibers in the striatum in a hemiparkinsonian rat model induced by unilateral injection of 6-hydroxydopamine (6-OHDA) [56]. Furthermore, a previous study performed in parkinsonian rats induced by the inhibition of the ubiquitin-proteasome system, another catabolic pathway different from autophagy, whose dysfunction also exerts a pathogenic role in PD, demonstrated that rTMS exerts neuroprotective effects by alleviating the loss of TH-positive dopaminergic neurons, by preventing the loss of striatal dopamine levels, by reducing the levels of apoptotic protein (cleaved caspase-3) and inflammatory factors (cyclooxygenase-2 and tumor necrosis factor alpha) in lesioned substantia nigra [57].

3. NIBS and Neuroprotection in Alzheimer’s Disease

Alzheimer’s disease (AD) is a neurodegenerative disorder clinically characterized by amnestic and non-amnestic cognitive impairments. In pathological neurons, β-amyloid (Aβ)-containing extracellular plaques and tau-containing intracellular neurofibrillary tangles determine the aggregation of misfolded proteins, which leads to microtubule disorganization, cholinergic dysfunction, neuroinflammation, OxS, and, ultimately, neural dysfunction and synaptic loss [58]. Current pharmacological treatments for AD are mostly symptomatic, and therapies altering the underlying pathological processes are not commonly available. Similar to PD, non-invasive brain stimulation (NIBS) techniques (e.g., tDCS, transcranial magnetic stimulation—TMS) were shown to improve AD symptoms (e.g., global cognition, cognitive and memory functions, executive performance) [59,60] (see Figure 2). Several randomized clinical trials have been conducted by using either tDCS or rTMS for the treatment of cognitive symptoms associated to AD [61,62,63]. However, to date, the biochemical mechanisms are still not fully understood (see Table 2). Neurotrophic factors (NTFs) regulate the growth, survival, proliferation, migration and differentiation of neurons [64] and have been extensively studied in the context of AD. In AD, the lowered expression of NTFs, such as nerve growth factor (NGF) [65], BDNF [66], glial cell line-derived neurotrophic factor (GDNF) [67] and ciliary neurotrophic factor (CNTF), have been observed in affected brain regions, including the temporal cortex and hippocampus [68]. Recently, particular interest has been aroused by the potentially beneficial effect of neuromodulation techniques on BDNF, which is required in the hippocampus for late-phase long-term potentiation and represents one of the most important cellular mechanisms that underlies learning and memory (see Table 2). Moreover, BDNF induces the secretion of acetylcholine by enhancing the differentiation and survival of cholinergic neurons in the basal forebrain [69]. Notably, various studies have recently shown increasing BDNF levels in the basal forebrain and hippocampus in AD animal models [70] as well as in the serum of AD patients [71] after rTMS when compared to controls. Similarly, rTMS was found to be effective on NGF brain levels [72,73]. Moreover, rTMS and tDCS were effective also on the BDNF-TrkB signalling pathway [38,55,74], which affects cell survival, migration, outgrowth of axons and dendrites, synaptogenesis, synaptic transmission, and synapse remodelling [75].
Beyond enhancing neuron survival, rTMS and tDCS concurrently inhibit apoptosis [76,77]. Specifically, TMS effectively balanced the apoptotic pathways in an AD-mice model [72], by inhibiting pro-apoptotic members of the Bcl-2 family, such as cleaved caspase-3 and Bax, which are usually overexpressed in AD. Conversely, the same study pointed out TMS-induced reduced ubiquitination of beta catenin, which notoriously promotes cell survival.
Brain tissue in AD patients is characterized by increased oxidative stress (OxS), due to an imbalance in Reactive Oxygen Species (ROS) and Reactive Nitrogen Species (RNS) levels and the antioxidant defense system, resulting in damage to proteins, lipids, and DNA oxidation/glycoxidation processes [78]. Velioglu and co-workers pointed out beneficial effects of rTMS on oxidative stress levels in AD patients by applying rTMS over the lateral parietal cortex [71]. In AD, OxS contributes to endothelial Nitric Oxide (NO) depletion [79,80] and quickening cognitive decline [81] due to cerebral hypoperfusion. Conversely, cerebral hypoperfusion is responsible for increased OxS. Trivedi et al. [82] applied low electrical fields to endothelial cells to induce the increase of NO levels, and, in turn, vasodilatation. Similarly, Marceglia et al. [83] speculated that tDCS may raise the NO level to prevent brain and hippocampal hypoperfusion. Further studies investigated NIBS’s beneficial effect on brain perfusion via blood-brain barrier modulation. Neurons, glial cells and cerebral blood vessels are closely related in structure and function, collectively referred to as the “neurovascular unit” (NVU), which is critical for regulating cerebral blood flow, maintaining both the blood-brain barrier integrity and signal transduction between cells. In AD brains, there is evidence that the cerebral microcirculation system is damaged and the main components of NVU underwent pathological changes [84]. tDCS has been demonstrated to decrease the number of glial cells and increase levels of vascular endothelial growth factor (VEGF) and interleukin-8, possibly resulting in decreased local inflammation, increased vascularization, improved toxic metabolite clearance and microcirculation protection [85,86,87]. A further study [88] showed that tDCS-treated AD model mice exhibited reduced glial fibrillary acidic protein (GFAP) levels. GFAP plays a crucial role in endothelial junction function and morphologic changes of astrocyte end foot processes [89]. NIBS might also prevent Aβ 1-42 aggregation, increasing Aβ serum levels and this was assessed by both tDCS in AD patients [90] and tACS in AD-mouse model [91], presumably via microglia activation [92].

4. Beyond Neurodegenerative Disorders: Neuroprotection and Stroke

Cerebrovascular disorders are far from the scope of our paper. Moreover, recent studies about neuroprotection and NIBS in stroke have been conducted on animal models only [93]. These studies suggest that NIBS, especially tDCS, have different effects on ischemia, which can be defined as “neuroprotective”, i.e., by modifying motor recovery over time; in particular, in rat models, non-invasive stimulation reduces blood-brain barrier (BBB) disruption [94], promotes microglia polarization [95] and decreases spreading depolarization [96]. However, these results have been obtained using different approaches and whether these effects lead to a clinical benefit is still a matter of debate. Moreover, it is worth noting that these models are based on an experimentally induced ischemia, so it is difficult to compare them with what happens in human beings.

5. Differences and Similarities in Neuroprotective Mechanisms between NIBS and DBS

Deep brain stimulation (DBS) consists in implantation of electrodes into specific regions of the brain for chronic transmission of continuous electrical stimulation in high frequency from an implantable pulse generator [97]. This neurosurgical model has become a well-established therapy in advanced stages of PD, in essential tremor and dystonia, with immediate beneficial effects on the motor symptoms [98]. New applications of DBS are now emerging in the field of other neurodegenerative diseases like AD and other neuropsychiatric conditions [22]. Besides clinical benefit, DBS has shown to chronically induce a reorganization of neural activity, influencing synaptic plasticity, neurotransmission and neurogenesis [99]. Also, some findings suggest a possible neuroprotective role [14], although the underlying mechanisms are still unclear. In this chapter, we focus on the possible neuroprotective effect of the DBS in neurodegenerative disease (i.e., PD and AD), and we compare the results with those from NIBS studies (see Figure 3).

5.1. DBS and NIBS: Neuroprotection in PD

It has been proposed that chronic subthalamic nucleus DBS (STN-DBS) might attenuate disease progression by preserving nigral dopamine neurons from degeneration [100,109,110,111], as reported for tDCS [17,44]. Although some results do not support this hypothesis [112,113], in MPTP-treated monkeys, the implantation of STN-DBS demonstrated to protect dopaminergic neuronal loss in the substantia nigra pars compacta (SNpc) and periaqueductal grey matter (PAG) [100,109]. Also, studies on 6-OHDA-treated rats, evaluated 5–7 days and 14 days after the 6-OHDA administration, found a significant increase in preservation of dopaminergic nigral neurons on the lesioned side [110,111]. Similar results, but with a different animal model (i.e., MPTP-treated mice [17] and 6-OHDA-treated mice [44]), were reached with anodal tDCS. Similarly, in line with tDCS results [17,44], Spieles-Engmann et al., 2020 [111] found that STN-DBS provided significant sparing of DA neurons in the SN of rats two and four weeks after 6-OHDA administration, when 50% of the DA cell loss had already occurred. This effect was dependent upon proper electrode placement within the STN [111].
The neuroprotective potential of STN-DBS was also studied in PD rat models undergoing genetic overexpression of viral vector wild-type alpha-synuclein, but with poor results [114]. Indeed, STN-DBS did not protect the striatal denervation and SNpc neuronal loss [114]. tDCS, at the cellular level, increases the release of neurotrophic factors (e.g., BNDF) [45]. The same effect was found for STN-DBS, which raised the level of neurotrophic factors (e.g., BNDF) in the nigrostriatal system and primary motor cortex [102]. The signalling created by BDNF binding with tropomyosin-related kinase type B (trkB), that is believed to occur at dendrites, triggers three different intracellular cascades: phospholipase Cgamma/protein kinase C (PLCgamma/PKC), which promotes the regulation of synaptic plasticity; mitogen-activated protein kinase/extracellular signal related-kinase (MAPK/ERK), which is involved in the protein synthesis; phosphatidilnositol 3-kinase/protein kinase B (PI3K/Akt), which inhibits apoptosis and regulates translation/trafficking. Regarding activation of this molecular pathway, STN-DBS was demonstrated to induce Akt phosphorylation in a study using 6-OHDA rat PD model [103]. A large number of glutamatergic efferents from the STN to the dopaminergic neurons in the SNpc becomes overactive in the abnormal parkinsonian basal ganglia network and has recently been implicated as an aetiological factor in the progressive decline of intact dopamine neurones in the SNc [115]. Some studies support the theory that the mechanism at the basis of the neuroprotection of the SNpc in STN-DBS is the reduction of overreactive glutamatergic projections originating from the STN [104,116]. Another study has recently shown that in a PD rat model, using viral vector-mediated nigrostriatal overexpression of human A53T alpha-synuclein, STN-DBS rescued tyrosine-hydroxylase (TH) expression in SNpc neurons and improved motor fluctuations, with no effect on loss of striatal dopamine levels [101]. Anodal tDCS demonstrated a close biochemical effect in MPTP-treated mice [19,45].

5.2. DBS and NIBS: Neuroprotection in AD

Recent investigations have discovered that NIBS and DBS might share neuroprotective mechanisms. For example, mice studies have demonstrated that DBS in the nucleus basalis of Meynert (NBM-DBS) increases the cholinergic transmission in the neocortex, inducing high secretion of NGF [105,106,117]. Similar effects were shown for NIBS techniques [72,73]. NGF is a factor that has an important role in the function and survival of cholinergic basal forebrain neurons [118] and in the improvement of synaptic plasticity [119]. From a histochemical point of view, NBM-DBS down-regulated A-beta40 and A-beta42, increasing the survival of neurons and reducing apoptotic cells in the hippocampus and cortex [106]. Also, by stimulating residual neurons, NBM-DBS reduced the progression of cognitive decline in patients in the early stages of disease [105,117]. The reduction of Aβ 1-42 aggregation is another mechanism of neuroprotection proposed for anodal tDCS [88,90], but also found during DBS [107,120]. For example, in 2019, Leplus et al. [120] showed, in implanted transgenic AD rat-model (TgF344-AD), a reduction of neuronal loss and amyloid burden in the hippocampus and cortex and a decrease of microglial proliferation, responsible for the altered clearance of Aβ. Also DBS in Enthorinal Cortex (EC-DBS) demonstrated improvement in cognitive domain, as spatial and contextual memory, in wild-type rat model [107], with a reduction in amyloid plaque burden [107]. Akwa et al. [108], using triple transgenic mouse model of AD, showed that EC-DBS rescued synaptophysin levels, which is implicated in synaptic transmission, counteracting the negative effects of tau phosphorylation in AD. This mechanism may involve the autophagic-lysosomal clearance of pathological forms of tau protein [121].
In contrast to NIBS [83,88], only few data support the role of DBS in the modulation of NO levels and, consequently, brain vasodilatation [122,123]. Phase I and II studies in patients with AD, using the Forniceal-DBS, have revealed a significant improvement of the metabolism in temporoparietal region, paralleled by a reduction of hippocampal atrophy, with no clinical improvements at one year [122,123].
Interestingly, growing evidence supporting a disease-modifying action of DBS comes from psychiatric disorders. Recent studies in animal models of schizophrenia have proved that DBS of the medial pre-frontal cortex (mPFC) can prevent the enlargement of lateral ventricle volumes, a marker of disease severity and progression, as well as the mal-development of both serotoninergic and glutamatergic transmission [124]. However, no corresponding considerations can be made for NIBS techniques.

6. Limitations and Future Directions

Despite the safety of these techniques, as demonstrated during years of clinical applications, studies regarding NIBS and neuroprotection have some limitations. First, in vitro and animal models do not completely fit with the complexity of human behavior. Also pathological bases may differ [125,126]; although animal models frequently express specific genes engaged in human neurodegenerative disorders, it is still debated whether there are intracellular aggregates of pathological proteins driving the disease course over time [127,128]. Another limitation is the deep localization of nuclei involved in the early phase of neurodegeneration; current modelling studies have explored the distribution of electric fields through the scalp, but little is known about the possibility to interfere with deepest structures, including basal ganglia and thalamus. An innovative approach has been recently provided by the use of the so-called non-invasive deep brain stimulation (NDBS), which is based on the collimation of different electric fields able to modulate deep nuclei in the brain and brainstem [129,130]; this method is safe, non-invasive and aims to assess if group-level hotspots exist in deep brain areas by using different electrode montages and has been successfully applied for the treatment of refractory temporal lobe epilepsies [131].
Finally, the same neurological disorder can be characterized by different pathological proteins, and, in some cases, pathological mechanisms are only poorly understood.

7. Conclusions

Neuroprotective and disease-modifying effects underlying the action of either invasive and non-invasive brain stimulating techniques represent a novel, promising and undervalued field of research. Recent evidence supports the role of tDCS in the modulation of the aggregation state and accumulation of pathological proteins. Nonetheless, only few studies have been recently developed using rTMS, whereas the literature describing a clear disease-modifying effect of DBS is scarce. DBS has recently shown neuroprotective effects in non-degenerative disorders, especially in psychiatric illnesses. tDCS seems to interfere both with lysosomal and non-lysosomal pathways, with the polarity of after-effects strictly depending on the aggregation state of pathological proteins and the pre-existing excitability state [34,132]. In this scenario, the classical, and highly debated, view supporting polarity-dependent after-effects, with cathodal tDCS inducing inhibition and anodal polarization leading to excitation [133], can be overcome. A third mechanism underlying a neuroprotective effect refers to the modulation of the inflammatory response, also by interfering with BBB permeability. This mechanism is also corroborated by findings in animal models of stroke [94]. Although α-synuclein and β-amyloid differ in terms of the number of aggregation states and intra-versus extra-cellular localization, encouraging results are coming from animal and cellular models of both PD and AD. Future studies should be devoted to evaluating the impact of NIBS in other “proteinopathies”, including Fronto-Temporal Dementias (FTDs) and Amyotrophic Lateral Sclerosis (ALS). Moreover, the neuroprotective role of so-called Non-Invasive Deep Brain Stimulation (NDBS) should be explored. NDBS refers to techniques currently under investigation that would allow to target deep brain regions through NIBS techniques [134]. Since tDCS and tACS are able to affect deep substrates primarily implicated in neurological pathologies (e.g., subthalamic nucleus for PD) [6], their neuroprotective action might be directed towards those structures, with huge clinical impact. However, no standardized and recognized stimulation protocols are now available for selective and precise targeting.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ferrucci, R.; Mameli, F.; Guidi, I.; Mrakic-Sposta, S.; Vergari, M.; Marceglia, S.; Cogiamanian, F.; Barbieri, S.; Scarpini, E.; Priori, A. Transcranial Direct Current Stimulation Improves Recognition Memory in Alzheimer Disease. Neurology 2008, 71, 493–498. [Google Scholar] [CrossRef] [Green Version]
  2. Lefaucheur, J.P.; Antal, A.; Ayache, S.S.; Benninger, D.H.; Brunelin, J.; Cogiamanian, F.; Cotelli, M.; De Ridder, D.; Ferrucci, R.; Langguth, B.; et al. Evidence-Based Guidelines on the Therapeutic Use of Transcranial Direct Current Stimulation (TDCS); Elsevier Ireland Ltd.: Amsterdam, The Netherlands, 2017; Volume 128, pp. 56–92. [Google Scholar]
  3. Cespón, J.; Rodella, C.; Miniussi, C.; Pellicciari, M.C. Behavioural and Electrophysiological Modulations Induced by Transcranial Direct Current Stimulation in Healthy Elderly and Alzheimer’s Disease Patients: A Pilot Study. Clin. Neurophysiol. 2019, 130, 2038–2052. [Google Scholar] [CrossRef] [PubMed]
  4. Chen, K.H.S.; Chen, R. Invasive and Noninvasive Brain Stimulation in Parkinson’s Disease: Clinical Effects and Future Perspectives. Clin. Pharmacol. Ther. 2019, 106, 763–775. [Google Scholar] [CrossRef] [PubMed]
  5. Ferrucci, R.; Bocci, T.; Cortese, F.; Ruggiero, F.; Priori, A. Cerebellar Transcranial Direct Current Stimulation in Neurological Disease; BioMed Central Ltd.: London, UK, 2016; Volume 3. [Google Scholar]
  6. Guidetti, M.; Arlotti, M.; Bocci, T.; Bianchi, A.M.; Parazzini, M.; Ferrucci, R.; Priori, A. Electric Fields Induced in the Brain by Transcranial Electric Stimulation: A Review of In Vivo Recordings. Biomedicines 2022, 10, 2333. [Google Scholar] [CrossRef] [PubMed]
  7. Liu, X.; Liu, H.; Liu, Z.; Rao, J.; Wang, J.; Wang, P.; Gong, X.; Wen, Y. Transcranial Direct Current Stimulation for Parkinson’s Disease: A Systematic Review and Meta-Analysis. Front. Aging Neurosci. 2021, 13, 691. [Google Scholar] [CrossRef]
  8. Dagan, M.; Herman, T.; Harrison, R.; Zhou, J.; Giladi, N.; Ruffini, G.; Manor, B.; Hausdorff, J.M. Multitarget Transcranial Direct Current Stimulation for Freezing of Gait in Parkinson’s Disease. Mov. Disord. 2018, 33, 642–646. [Google Scholar] [CrossRef] [PubMed]
  9. Manor, B.; Dagan, M.; Herman, T.; Gouskova, N.A.; Vanderhorst, V.G.; Giladi, N.; Travison, T.G.; Pascual-Leone, A.; Lipsitz, L.A.; Hausdorff, J.M. Multitarget Transcranial Electrical Stimulation for Freezing of Gait: A Randomized Controlled Trial. Mov. Disord. 2021, 36, 2693–2698. [Google Scholar] [CrossRef]
  10. Putzolu, M.; Ogliastro, C.; Lagravinese, G.; Bonassi, G.; Trompetto, C.; Marchese, R.; Avanzino, L.; Pelosin, E. Investigating the Effects of Transcranial Direct Current Stimulation on Obstacle Negotiation Performance in Parkinson Disease with Freezing of Gait: A Pilot Study. Brain Stimul. 2019, 12, 1583–1585. [Google Scholar] [CrossRef]
  11. Valentino, F.; Cosentino, G.; Brighina, F.; Pozzi, N.G.; Sandrini, G.; Fierro, B.; Savettieri, G.; D’Amelio, M.; Pacchetti, C. Transcranial Direct Current Stimulation for Treatment of Freezing of Gait: A Cross-over Study. Mov. Disord. 2014, 29, 1064–1069. [Google Scholar] [CrossRef]
  12. Rascol, O.; Payoux, P.; Ory, F.; Ferreira, J.J.; Brefel-Courbon, C.; Montastruc, J.-L. Limitations of Current Parkinson’s Disease Therapy. Ann. Neurol. 2003, 53 (Suppl. S3), S3–S15. [Google Scholar] [CrossRef]
  13. Hadar, R.; Winter, R.; Edemann-Callesen, H.; Wieske, F.; Habelt, B.; Khadka, N.; Felgel-Farnholz, V.; Barroeta-Hlusicka, E.; Reis, J.; Tatarau, C.A.; et al. Prevention of Schizophrenia Deficits via Non-Invasive Adolescent Frontal Cortex Stimulation in Rats. Mol. Psychiatry 2019, 25, 896–905. [Google Scholar] [CrossRef] [PubMed]
  14. McKinnon, C.; Gros, P.; Lee, D.J.; Hamani, C.; Lozano, A.M.; Kalia, L.V.; Kalia, S.K. Deep Brain Stimulation: Potential for Neuroprotection. Ann. Clin. Transl. Neurol. 2018, 6, 174–185. [Google Scholar] [CrossRef] [Green Version]
  15. de Lau, L.M.; Breteler, M.M. Epidemiology of Parkinson’s Disease. Lancet Neurol. 2006, 5, 525–535. [Google Scholar] [CrossRef]
  16. Balestrino, R.; Schapira, A.H.V. Parkinson Disease. Eur. J. Neurol. 2020, 27, 27–42. [Google Scholar] [CrossRef]
  17. Lee, S.B.; Youn, J.; Jang, W.; Yang, H.O. Neuroprotective Effect of Anodal Transcranial Direct Current Stimulation on 1-Methyl-4-Phenyl-1,2,3,6-Tetrahydropyridine (MPTP)-Induced Neurotoxicity in Mice through Modulating Mitochondrial Dynamics. Neurochem. Int. 2019, 129, 104491. [Google Scholar] [CrossRef]
  18. Hattori, N.; Mizuno, Y. Mitochondrial Dysfunction in Parkinson’s Disease. Exp. Neurobiol. 2015, 24, 103. [Google Scholar] [CrossRef] [Green Version]
  19. Li, X.; Lu, C.; Wei, Y.; Hu, R.; Wang, Y.; Li, K. Transcranial Direct Current Stimulation Ameliorates Behavioral Deficits and Reduces Oxidative Stress in 1-Methyl-4-Phenyl-1,2,3,6-Tetrahydropyridine-Induced Mouse Model of Parkinson’s Disease. Neuromodul. Technol. Neural Interface 2015, 18, 442–447. [Google Scholar] [CrossRef]
  20. Laste, G.; Caumo, W.; Adachi, L.N.S.; Rozisky, J.R.; De MacEdo, I.C.; Filho, P.R.M.; Partata, W.A.; Fregni, F.; Torres, I.L.S. After-Effects of Consecutive Sessions of Transcranial Direct Current Stimulation (TDCS) in a Rat Model of Chronic Inflammation. Exp. Brain Res. 2012, 221, 75–83. [Google Scholar] [CrossRef]
  21. Lynch-Day, M.A.; Mao, K.; Wang, K.; Zhao, M.; Klionsky, D.J. The Role of Autophagy in Parkinson’s Disease. Cold Spring Harb. Perspect. Med. 2012, 2, a009357. [Google Scholar] [CrossRef] [Green Version]
  22. Guidetti, M.; Marceglia, S.; Loh, A.; Harmsen, I.E.; Meoni, S.; Foffani, G.; Lozano, A.M.; Moro, E.; Volkmann, J.; Priori, A. Clinical Perspectives of Adaptive Deep Brain Stimulation. Brain Stimul. 2021, 14, 1238–1247. [Google Scholar] [CrossRef]
  23. Hariz, M.; Blomstedt, P. Deep Brain Stimulation for Parkinson’s Disease. J. Intern. Med. 2022, 292, 764–778. [Google Scholar] [CrossRef] [PubMed]
  24. 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]
  25. Madrid, J.; Benninger, D.H. Non-Invasive Brain Stimulation for Parkinson’s Disease: Clinical Evidence, Latest Concepts and Future Goals: A Systematic Review. J. Neurosci. Methods 2021, 347, 108957. [Google Scholar] [CrossRef]
  26. Ganguly, J.; Murgai, A.; Sharma, S.; Aur, D.; Jog, M. Non-Invasive Transcranial Electrical Stimulation in Movement Disorders. Front. Neurosci. 2020, 14, 522. [Google Scholar] [CrossRef] [PubMed]
  27. 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 2010, 81, 1105–1111. [Google Scholar] [CrossRef] [Green Version]
  28. Simpson, M.W.; Mak, M. The Effect of Transcranial Direct Current Stimulation on Upper Limb Motor Performance in Parkinson’s Disease: A Systematic Review. J. Neurol. 2020, 267, 3479–3488. [Google Scholar] [CrossRef]
  29. Broeder, S.; Nackaerts, E.; Heremans, E.; Vervoort, G.; Meesen, R.; Verheyden, G.; Nieuwboer, A. Transcranial Direct Current Stimulation in Parkinson’s Disease: Neurophysiological Mechanisms and Behavioral Effects. Neurosci. Biobehav. Rev. 2015, 57, 105–117. [Google Scholar] [CrossRef]
  30. Fregni, F.; Boggio, P.S.; Santos, M.C.; Lima, M.; Vieira, A.L.; Rigonatti, S.P.; Silva, M.T.A.; Barbosa, E.R.; Nitsche, M.A.; Pascual-Leone, A. Noninvasive Cortical Stimulation with Transcranial Direct Current Stimulation in Parkinson’s Disease. Mov. Disord. 2006, 21, 1693–1702. [Google Scholar] [CrossRef]
  31. Elsner, B.; Kugler, J.; Pohl, M.; Mehrholz, J. Transcranial Direct Current Stimulation (TDCS) for Idiopathic Parkinson’s Disease. Cochrane Database Syst. Rev. 2016, 2016, CD010916. [Google Scholar] [CrossRef]
  32. Doruk, D.; Gray, Z.; Bravo, G.L.; Pascual-Leone, A.; Fregni, F. Effects of TDCS on Executive Function in Parkinson’s Disease. Neurosci. Lett. 2014, 582, 27–31. [Google Scholar] [CrossRef]
  33. Manto, M.; Argyropoulos, G.P.D.; Bocci, T.; Celnik, P.A.; Corben, L.A.; Guidetti, M.; Koch, G.; Priori, A.; Rothwell, J.C.; Sadnicka, A.; et al. Consensus Paper: Novel Directions and Next Steps of Non-Invasive Brain Stimulation of the Cerebellum in Health and Disease. Cerebellum 2021, 21, 1092–1122. [Google Scholar] [CrossRef] [PubMed]
  34. Sala, G.; Bocci, T.; Borzì, V.; Parazzini, M.; Priori, A.; Ferrarese, C. Direct Current Stimulation Enhances Neuronal Alpha-Synuclein Degradation in Vitro. Sci. Rep. 2021, 11, 2197. [Google Scholar] [CrossRef]
  35. Fukai, M.; Bunai, T.; Hirosawa, T.; Kikuchi, M.; Ito, S.; Minabe, Y.; Ouchi, Y. Endogenous Dopamine Release under Transcranial Direct-Current Stimulation Governs Enhanced Attention: A Study with Positron Emission Tomography. Transl. Psychiatry 2019, 9, 115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Fonteneau, C.; Redoute, J.; Haesebaert, F.; Le Bars, D.; Costes, N.; Suaud-Chagny, M.F.; Brunelin, J. Frontal Transcranial Direct Current Stimulation Induces Dopamine Release in the Ventral Striatum in Human. Cereb. Cortex 2018, 28, 2636–2646. [Google Scholar] [CrossRef]
  37. Tanaka, T.; Takano, Y.; Tanaka, S.; Hironaka, N.; Kobayashi, K.; Hanakawa, T.; Watanabe, K.; Honda, M. Transcranial Direct-Current Stimulation Increases Extracellular Dopamine Levels in the Rat Striatum. Front. Syst. Neurosci. 2013, 7, 6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. 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]
  39. Spezia Adachi, L.N.; Caumo, W.; Laste, G.; Fernandes Medeiros, L.; Ripoll Rozisky, J.; De Souza, A.; Fregni, F.; Torres, I.L.S. Reversal of Chronic Stress-Induced Pain by Transcranial Direct Current Stimulation (TDCS) in an Animal Model. Brain Res. 2012, 1489, 17–26. [Google Scholar] [CrossRef] [Green Version]
  40. Ranieri, F.; Podda, M.V.; Riccardi, E.; Frisullo, G.; Dileone, M.; Profice, P.; Pilato, F.; di Lazzaro, V.; Grassi, C. Modulation of LTP at Rat Hippocampal CA3-CA1 Synapses by Direct Current Stimulation. J. Neurophysiol. 2012, 107, 1868–1880. [Google Scholar] [CrossRef] [Green Version]
  41. Zigova, T.; Pencea, V.; Wiegand, S.J.; Luskin, M.B. Intraventricular Administration of BDNF Increases the Number of Newly Generated Neurons in the Adult Olfactory Bulb. Mol. Cell. Neurosci. 1998, 11, 234–245. [Google Scholar] [CrossRef]
  42. Benraiss, A.; Chmielnicki, E.; Lerner, K.; Roh, D.; Goldman, S.A. Adenoviral Brain-Derived Neurotrophic Factor Induces Both Neostriatal and Olfactory Neuronal Recruitment from Endogenous Progenitor Cells in the Adult Forebrain. J. Neurosci. 2001, 21, 6718–6731. [Google Scholar] [CrossRef]
  43. Huang, E.J.; Reichardt, L.F. Neurotrophins: Roles in Neuronal Development and Function. Annu. Rev. Neurosci. 2001, 24, 677. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Feng, X.J.; Huang, Y.T.; Huang, Y.Z.; Kuo, C.W.; Peng, C.W.; Rotenberg, A.; Juan, C.H.; Pei, Y.C.; Chen, Y.H.; Chen, K.Y.; et al. Early Transcranial Direct Current Stimulation Treatment Exerts Neuroprotective Effects on 6-OHDA-Induced Parkinsonism in Rats. Brain Stimul. 2020, 13, 655–663. [Google Scholar] [CrossRef] [Green Version]
  45. Lee, S.B.; Kim, H.T.; Yang, H.O.; Jang, W. Anodal Transcranial Direct Current Stimulation Prevents Methyl-4-Phenyl-1,2,3,6-Tetrahydropyridine (MPTP)-Induced Neurotoxicity by Modulating Autophagy in an in Vivo Mouse Model of Parkinson’s Disease. Sci. Rep. 2018, 8, 15165. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Alam, G.; Richardson, J.R. Regulation of Tyrosine Hydroxylase: Relevance to Parkinson’s Disease. Genet. Neurol. Behav. Diet Park. Dis. 2020, 2, 51–66. [Google Scholar] [CrossRef]
  47. Ischiropoulos, H.; Beckman, J.S. Oxidative Stress and Nitration in Neurodegeneration: Cause, Effect, or Association? J. Clin. Investig. 2003, 111, 163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Deas, E.; Wood, N.W.; Plun-Favreau, H. Mitophagy and Parkinson’s Disease: The PINK1–Parkin Link. Biochim. Biophys. Acta-Mol. Cell Res. 2011, 1813, 623–633. [Google Scholar] [CrossRef] [Green Version]
  49. Rae, C.D.; Lee, V.H.C.; Ordidge, R.J.; Alonzo, A.; Loo, C. Anodal Transcranial Direct Current Stimulation Increases Brain Intracellular PH and Modulates Bioenergetics. Int. J. Neuropsychopharmacol. 2013, 16, 1695–1706. [Google Scholar] [CrossRef] [Green Version]
  50. Filichia, E.; Hoffer, B.; Qi, X.; Luo, Y. Inhibition of Drp1 Mitochondrial Translocation Provides Neural Protection in Dopaminergic System in a Parkinson’s Disease Model Induced by MPTP. Sci. Rep. 2016, 6, 32656. [Google Scholar] [CrossRef] [Green Version]
  51. Majeski, A.E.; Fred Dice, J. Mechanisms of Chaperone-Mediated Autophagy. Int. J. Biochem. Cell Biol. 2004, 36, 2435–2444. [Google Scholar] [CrossRef]
  52. Cuervo, A.M.; Stafanis, L.; Fredenburg, R.; Lansbury, P.T.; Sulzer, D. Impaired Degradation of Mutant Alpha-Synuclein by Chaperone-Mediated Autophagy. Science 2004, 305, 1292–1295. [Google Scholar] [CrossRef]
  53. Lu, J.; Wu, M.; Yue, Z. Autophagy and Parkinson’s Disease. Adv. Exp. Med. Biol. 2020, 1207, 21–51. [Google Scholar] [CrossRef] [PubMed]
  54. Vekrellis, K.; Stefanis, L. Targeting Intracellular and Extracellular Alpha-Synuclein as a Therapeutic Strategy in Parkinson’s Disease and Other Synucleinopathies. Expert Opin. Ther. Targets 2012, 16, 421–432. [Google Scholar] [CrossRef] [PubMed]
  55. Chen, A.; Xiong, L.J.; Tong, Y.; Mao, M. Neuroprotective Effect of Brain-Derived Neurotrophic Factor Mediated by Autophagy through the PI3K/Akt/MTOR Pathway. Mol. Med. Rep. 2013, 8, 1011–1016. [Google Scholar] [CrossRef] [Green Version]
  56. Hsieh, T.H.; He, X.K.; Liu, H.H.; Chen, J.J.J.; Peng, C.W.; Liu, H.L.; Rotenberg, A.; Chen, K.T.; Chang, M.Y.; Chiang, Y.H.; et al. Early Repetitive Transcranial Magnetic Stimulation Exerts Neuroprotective Effects and Improves Motor Functions in Hemiparkinsonian Rats. Neural Plast. 2021, 2021, 1763533. [Google Scholar] [CrossRef] [PubMed]
  57. Ba, M.; Ma, G.; Ren, C.; Sun, X.; Kong, M. Repetitive Transcranial Magnetic Stimulation for Treatment of Lactacystin-Induced Parkinsonian Rat Model. Oncotarget 2017, 8, 50921–50929. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Knopman, D.S.; Amieva, H.; Petersen, R.C.; Chételat, G.; Holtzman, D.M.; Hyman, B.T.; Nixon, R.A.; Jones, D.T. Alzheimer Disease. Nat. Rev. Dis. Prim. 2021, 7, 33. [Google Scholar] [CrossRef]
  59. Chou, Y.H.; Ton That, V.; Sundman, M. A Systematic Review and Meta-Analysis of RTMS Effects on Cognitive Enhancement in Mild Cognitive Impairment and Alzheimer’s Disease. Neurobiol. Aging 2020, 86, 1–10. [Google Scholar] [CrossRef]
  60. Cammisuli, D.M.; Cignoni, F.; Ceravolo, R.; Bonuccelli, U.; Castelnuovo, G. Transcranial Direct Current Stimulation (TDCS) as a Useful Rehabilitation Strategy to Improve Cognition in Patients With Alzheimer’s Disease and Parkinson’s Disease: An Updated Systematic Review of Randomized Controlled Trials. Front. Neurol. 2022, 12, 2648. [Google Scholar] [CrossRef]
  61. Rajji, T.K. Transcranial Magnetic and Electrical Stimulation in Alzheimer’s Disease and Mild Cognitive Impairment: A Review of Randomized Controlled Trials. Clin. Pharmacol. Ther. 2019, 106, 776–780. [Google Scholar] [CrossRef]
  62. Teselink, J.; Bawa, K.K.; Koo, G.K.; Sankhe, K.; Liu, C.S.; Rapoport, M.; Oh, P.; Marzolini, S.; Gallagher, D.; Swardfager, W.; et al. Efficacy of Non-Invasive Brain Stimulation on Global Cognition and Neuropsychiatric Symptoms in Alzheimer’s Disease and Mild Cognitive Impairment: A Meta-Analysis and Systematic Review. Ageing Res. Rev. 2021, 72, 101499. [Google Scholar] [CrossRef]
  63. Lefaucheur, J.P. A Comprehensive Database of Published TDCS Clinical Trials (2005–2016). Neurophysiol. Clin. 2016, 46, 319–398. [Google Scholar] [CrossRef] [PubMed]
  64. Xiao, N.; Le, Q.T. Neurotrophic Factors and Their Potential Applications in Tissue Regeneration. Arch. Immunol. Ther. Exp. (Warsz) 2016, 64, 89–99. [Google Scholar] [CrossRef]
  65. Mitra, S.; Behbahani, H.; Eriksdotter, M. Innovative Therapy for Alzheimer’s Disease-With Focus on Biodelivery of NGF. Front. Neurosci. 2019, 13, 38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Ng, T.K.S.; Ho, C.S.H.; Tam, W.W.S.; Kua, E.H.; Ho, R.C.M. Decreased Serum Brain-Derived Neurotrophic Factor (BDNF) Levels in Patients with Alzheimer’s Disease (AD): A Systematic Review and Meta-Analysis. Int. J. Mol. Sci. 2019, 20, 257. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Straten, G.; Eschweiler, G.W.; Maetzler, W.; Laske, C.; Leyhe, T. Glial Cell-Line Derived Neurotrophic Factor (GDNF) Concentrations in Cerebrospinal Fluid and Serum of Patients with Early Alzheimer’s Disease and Normal Controls. J. Alzheimer’s Dis. 2009, 18, 331–337. [Google Scholar] [CrossRef]
  68. Rockenstein, E.; Ubhi, K.; Doppler, E.; Novak, P.; Moessler, H.; Li, B.; Blanchard, J.; Grundke-Iqbal, I.; Iqbal, K.; Mante, M.; et al. Regional Comparison of the Neurogenic Effects of CNTF-Derived Peptides and Cerebrolysin in AβPP Transgenic Mice. J. Alzheimer’s Dis. 2011, 27, 743–752. [Google Scholar] [CrossRef] [Green Version]
  69. Knipper, M.; da Penha Berzaghi, M.; Blöchl, A.; Breer, H.; Thoenen, H.; Lindholm, D. Positive Feedback between Acetylcholine and the Neurotrophins Nerve Growth Factor and Brain-Derived Neurotrophic Factor in the Rat Hippocampus. Eur. J. Neurosci. 1994, 6, 668–671. [Google Scholar] [CrossRef]
  70. Choung, J.S.; Kim, J.M.; Ko, M.H.; Cho, D.S.; Kim, M.Y. Therapeutic Efficacy of Repetitive Transcranial Magnetic Stimulation in an Animal Model of Alzheimer’s Disease. Sci. Rep. 2021, 11, 437. [Google Scholar] [CrossRef]
  71. Velioglu, H.A.; Hanoglu, L.; Bayraktaroglu, Z.; Toprak, G.; Guler, E.M.; Bektay, M.Y.; Mutlu-Burnaz, O.; Yulug, B. Left Lateral Parietal RTMS Improves Cognition and Modulates Resting Brain Connectivity in Patients with Alzheimer’s Disease: Possible Role of BDNF and Oxidative Stress. Neurobiol. Learn. Mem. 2021, 180, 107410. [Google Scholar] [CrossRef]
  72. Chen, X.; Chen, S.; Liang, W.; Ba, F. Administration of Repetitive Transcranial Magnetic Stimulation Attenuates A β 1-42-Induced Alzheimer’s Disease in Mice by Activating β-Catenin Signaling. Biomed Res. Int. 2019, 2019, 1431760. [Google Scholar] [CrossRef]
  73. Tan, T.; Xie, J.; Liu, T.; Chen, X.; Zheng, X.; Tong, Z.; Tian, X. Low-Frequency (1Hz) Repetitive Transcranial Magnetic Stimulation (RTMS) Reverses Aβ1–42-Mediated Memory Deficits in Rats. Exp. Gerontol. 2013, 48, 786–794. [Google Scholar] [CrossRef] [PubMed]
  74. Chen, X.; Dong, G.Y.; Wang, L.X. High-Frequency Transcranial Magnetic Stimulation Protects APP/PS1 Mice against Alzheimer’s Disease Progress by Reducing APOE and Enhancing Autophagy. Brain Behav. 2020, 10, e01740. [Google Scholar] [CrossRef] [PubMed]
  75. Ohira, K.; Hayashi, M. A New Aspect of the TrkB Signaling Pathway in Neural Plasticity. Curr. Neuropharmacol. 2009, 7, 276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Obulesu, M.; Lakshmi, M.J. Apoptosis in Alzheimer’s Disease: An Understanding of the Physiology, Pathology and Therapeutic Avenues. Neurochem. Res. 2014, 39, 2301–2312. [Google Scholar] [CrossRef]
  77. Paradis, E.; Douillard, H.; Koutroumanis, M.; Goodyer, C.; LeBlanc, A. Amyloid Beta Peptide of Alzheimer’s Disease Downregulates Bcl-2 and Upregulates Bax Expression in Human Neurons. J. Neurosci. 1996, 16, 7533–7539. [Google Scholar] [CrossRef] [Green Version]
  78. Feng, Y.; Wang, X. Antioxidant Therapies for Alzheimer’s Disease. Oxid. Med. Cell. Longev. 2012, 2012, 472932. [Google Scholar] [CrossRef] [Green Version]
  79. Selley, M.L. Increased Concentrations of Homocysteine and Asymmetric Dimethylarginine and Decreased Concentrations of Nitric Oxide in the Plasma of Patients with Alzheimer’s Disease. Neurobiol. Aging 2003, 24, 903–907. [Google Scholar] [CrossRef]
  80. Guix, F.X.; Uribesalgo, I.; Coma, M.; Muñoz, F.J. The Physiology and Pathophysiology of Nitric Oxide in the Brain. Prog. Neurobiol. 2005, 76, 126–152. [Google Scholar] [CrossRef]
  81. Katusic, Z.S.; Austin, S.A. Endothelial Nitric Oxide: Protector of a Healthy Mind. Eur. Heart J. 2014, 35, 888–894. [Google Scholar] [CrossRef] [Green Version]
  82. Trivedi, D.P.; Hallock, K.J.; Bergethon, P.R. Electric Fields Caused by Blood Flow Modulate Vascular Endothelial Electrophysiology and Nitric Oxide Production. Bioelectromagnetics 2013, 34, 22–30. [Google Scholar] [CrossRef]
  83. Marceglia, S.; Mrakic-Sposta, S.; Rosa, M.; Ferrucci, R.; Mameli, F.; Vergari, M.; Arlotti, M.; Ruggiero, F.; Scarpini, E.; Galimberti, D.; et al. Transcranial Direct Current Stimulation Modulates Cortical Neuronal Activity in Alzheimer’s Disease. Front. Neurosci. 2016, 10, 134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Kirabali, T.; Rust, R.; Rigotti, S.; Siccoli, A.; Nitsch, R.M.; Kulic, L. Distinct Changes in All Major Components of the Neurovascular Unit across Different Neuropathological Stages of Alzheimer’s Disease. Brain Pathol. 2020, 30, 1056–1070. [Google Scholar] [CrossRef] [PubMed]
  85. Bai, H.; Forrester, J.V.; Zhao, M. DC Electric Stimulation Upregulates Angiogenic Factors in Endothelial Cells through Activation of VEGF Receptors. Cytokine 2011, 55, 110–115. [Google Scholar] [CrossRef] [Green Version]
  86. Rueger, M.A.; Keuters, M.H.; Walberer, M.; Braun, R.; Klein, R.; Sparing, R.; Fink, G.R.; Graf, R.; Schroeter, M. Multi-Session Transcranial Direct Current Stimulation (TDCS) Elicits Inflammatory and Regenerative Processes in the Rat Brain. PLoS ONE 2012, 7, e43776. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Cancel, L.M.; Arias, K.; Bikson, M.; Tarbell, J.M. Direct Current Stimulation of Endothelial Monolayers Induces a Transient and Reversible Increase in Transport Due to the Electroosmotic Effect. Sci. Rep. 2018, 8, 9265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Luo, Y.; Yang, H.; Yan, X.; Wu, Y.; Wei, G.; Wu, X.; Tian, X.; Xiong, Y.; Wu, G.; Wen, H. Transcranial Direct Current Stimulation Alleviates Neurovascular Unit Dysfunction in Mice With Preclinical Alzheimer’s Disease. Front. Aging Neurosci. 2022, 14, 857415. [Google Scholar] [CrossRef]
  89. Teter, B. Rodent Aging. In Encyclopedia of Neuroscience; Academic Press: Oxford, UK, 2009; pp. 397–406. [Google Scholar] [CrossRef]
  90. Khedr, E.M.; Salama, R.H.; Abdel Hameed, M.; Abo Elfetoh, N.; Seif, P. Therapeutic Role of Transcranial Direct Current Stimulation in Alzheimer Disease Patients: Double-Blind, Placebo-Controlled Clinical Trial. Neurorehabil. Neural Repair 2019, 33, 384–394. [Google Scholar] [CrossRef]
  91. Dhaynaut, M.; Sprugnoli, G.; Cappon, D.; Macone, J.; Sanchez, J.S.; Normandin, M.D.; Guehl, N.J.; Koch, G.; Paciorek, R.; Connor, A.; et al. Impact of 40 Hz Transcranial Alternating Current Stimulation on Cerebral Tau Burden in Patients with Alzheimer’s Disease: A Case Series. J. Alzheimer’s Dis. 2022, 85, 1667–1676. [Google Scholar] [CrossRef]
  92. Luo, Y.; Yang, W.; Li, N.; Yang, X.; Zhu, B.; Wang, C.; Hou, W.; Wang, X.; Wen, H.; Tian, X. Anodal Transcranial Direct Current Stimulation Can Improve Spatial Learning and Memory and Attenuate Aβ 42 Burden at the Early Stage of Alzheimer’s Disease in APP/PS1 Transgenic Mice. Front. Aging Neurosci. 2020, 12, 134. [Google Scholar] [CrossRef]
  93. Storch, S.; Samantzis, M.; Balbi, M. Driving Oscillatory Dynamics: Neuromodulation for Recovery After Stroke. Front. Syst. Neurosci. 2021, 15, 65. [Google Scholar] [CrossRef]
  94. Yang, Y.; Yang, L.Y.; Orban, L.; Cuylear, D.; Thompson, J.; Simon, B.; Yang, Y. Non-Invasive Vagus Nerve Stimulation Reduces Blood-Brain Barrier Disruption in a Rat Model of Ischemic Stroke. Brain Stimul. 2018, 11, 689–698. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Zhao, X.P.; Zhao, Y.; Qin, X.Y.; Wan, L.Y.; Fan, X.X. Non-Invasive Vagus Nerve Stimulation Protects Against Cerebral Ischemia/Reperfusion Injury and Promotes Microglial M2 Polarization Via Interleukin-17A Inhibition. J. Mol. Neurosci. 2019, 67, 217–226. [Google Scholar] [CrossRef] [PubMed]
  96. Notturno, F.; Pace, M.; Zappasodi, F.; Cam, E.; Bassetti, C.L.; Uncini, A. Neuroprotective Effect of Cathodal Transcranial Direct Current Stimulation in a Rat Stroke Model. J. Neurol. Sci. 2014, 342, 146–151. [Google Scholar] [CrossRef]
  97. Priori, A.; Maiorana, N.; Dini, M.; Guidetti, M.; Marceglia, S.; Ferrucci, R. Adaptive Deep Brain Stimulation (ADBS). Int. Rev. Neurobiol. 2021, 159, 111–127. [Google Scholar] [CrossRef] [PubMed]
  98. Lozano, A.M.; Lipsman, N.; Bergman, H.; Brown, P.; Chabardes, S.; Chang, J.W.; Matthews, K.; McIntyre, C.C.; Schlaepfer, T.E.; Schulder, M.; et al. Deep Brain Stimulation: Current Challenges and Future Directions. Nat. Rev. Neurol. 2019, 15, 148–160. [Google Scholar] [CrossRef] [PubMed]
  99. Herrington, T.M.; Cheng, J.J.; Eskandar, E.N. Mechanisms of Deep Brain Stimulation. J. Neurophysiol. 2016, 115, 19–38. [Google Scholar] [CrossRef] [Green Version]
  100. Wallace, B.A.; Ashkan, K.; Heise, C.E.; Foote, K.D.; Torres, N.; Mitrofanis, J.; Benabid, A.L. Survival of Midbrain Dopaminergic Cells after Lesion or Deep Brain Stimulation of the Subthalamic Nucleus in MPTP-Treated Monkeys. Brain 2007, 130, 2129–2145. [Google Scholar] [CrossRef] [Green Version]
  101. Musacchio, T.; Rebenstorff, M.; Fluri, F.; Brotchie, J.M.; Volkmann, J.; Koprich, J.B.; Ip, C.W. Subthalamic Nucleus Deep Brain Stimulation Is Neuroprotective in the A53T α-Synuclein Parkinson’s Disease Rat Model. Ann. Neurol. 2017, 81, 825–836. [Google Scholar] [CrossRef] [Green Version]
  102. Spieles-Engemann, A.L.; Steece-Collier, K.; Behbehani, M.M.; Collier, T.J.; Wohlgenant, S.L.; Kemp, C.J.; Cole-Strauss, A.; Levine, N.D.; Gombash, S.E.; Thompson, V.B.; et al. Subthalamic Nucleus Stimulation Increases Brain Derived Neurotrophic Factor in the Nigrostriatal System and Primary Motor Cortex. J. Park. Dis. 2011, 1, 123. [Google Scholar] [CrossRef]
  103. Fischer, D.L.; Kemp, C.J.; Cole-Strauss, A.; Polinski, N.K.; Paumier, K.L.; Lipton, J.W.; Steece-Collier, K.; Collier, T.J.; Buhlinger, D.J.; Sortwell, C.E. Subthalamic Nucleus Deep Brain Stimulation Employs TrkB Signaling for Neuroprotection and Functional Restoration. J. Neurosci. 2017, 37, 6786–6796. [Google Scholar] [CrossRef] [Green Version]
  104. Rodriguez, M.C.; Obeso, J.A.; Olanow, C.W. Subthalamic Nucleus-Mediated Excitotoxicity in Parkinson’s Disease: A Target for Neuroprotection. Ann. Neurol. 1998, 44, S175–S188. [Google Scholar] [CrossRef] [PubMed]
  105. Baldermann, J.C.; Hardenacke, K.; Hu, X.; Köster, P.; Horn, A.; Freund, H.J.; Zilles, K.; Sturm, V.; Visser-Vandewalle, V.; Jessen, F.; et al. Neuroanatomical Characteristics Associated with Response to Deep Brain Stimulation of the Nucleus Basalis of Meynert for Alzheimer’s Disease. Neuromodulation 2018, 21, 184–190. [Google Scholar] [CrossRef] [PubMed]
  106. Huang, C.; Chu, H.; Ma, Y.; Zhou, Z.; Dai, C.; Huang, X.; Fang, L.; Ao, Q.; Huang, D. The Neuroprotective Effect of Deep Brain Stimulation at Nucleus Basalis of Meynert in Transgenic Mice with Alzheimer’s Disease. Brain Stimul. 2019, 12, 161–174. [Google Scholar] [CrossRef] [PubMed]
  107. Xia, F.; Yiu, A.; Stone, S.S.D.; Oh, S.; Lozano, A.M.; Josselyn, S.A.; Frankland, P.W. Entorhinal Cortical Deep Brain Stimulation Rescues Memory Deficits in Both Young and Old Mice Genetically Engineered to Model Alzheimer’s Disease. Neuropsychopharmacology 2017, 42, 2493–2503. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Akwa, Y.; Gondard, E.; Mann, A.; Capetillo-Zarate, E.; Alberdi, E.; Matute, C.; Marty, S.; Vaccari, T.; Lozano, A.M.; Baulieu, E.E.; et al. Synaptic Activity Protects against AD and FTD-like Pathology via Autophagic-Lysosomal Degradation. Mol. Psychiatry 2017, 23, 1530–1540. [Google Scholar] [CrossRef] [Green Version]
  109. Shaw, V.E.; Keay, K.A.; Ashkan, K.; Benabid, A.L.; Mitrofanis, J. Dopaminergic Cells in the Periaqueductal Grey Matter of MPTP-Treated Monkeys and Mice; Patterns of Survival and Effect of Deep Brain Stimulation and Lesion of the Subthalamic Nucleus. Park. Relat. Disord. 2010, 16, 338–344. [Google Scholar] [CrossRef]
  110. Harnack, D.; Meissner, W.; Jira, J.A.; Winter, C.; Morgenstern, R.; Kupsch, A. Placebo-Controlled Chronic High-Frequency Stimulation of the Subthalamic Nucleus Preserves Dopaminergic Nigral Neurons in a Rat Model of Progressive Parkinsonism. Exp. Neurol. 2008, 210, 257–260. [Google Scholar] [CrossRef]
  111. Spieles-Engemann, A.L.; Behbehani, M.M.; Collier, T.J.; Wohlgenant, S.L.; Steece-Collier, K.; Paumier, K.; Daley, B.F.; Gombash, S.; Madhavan, L.; Mandybur, G.T.; et al. Stimulation of the Rat Subthalamic Nucleus Is Neuroprotective Following Significant Nigral Dopamine Neuron Loss. Neurobiol. Dis. 2010, 39, 105–115. [Google Scholar] [CrossRef] [Green Version]
  112. Hilker, R.; Portman, A.T.; Voges, J.; Staal, M.J.; Burghaus, L.; Van Laar, T.; Koulousakis, A.; Maguire, R.P.; Pruim, J.; De Jong, B.M.; et al. Disease Progression Continues in Patients with Advanced Parkinson’s Disease and Effective Subthalamic Nucleus Stimulation. J. Neurol. Neurosurg. Psychiatry 2005, 76, 1217–1221. [Google Scholar] [CrossRef]
  113. Pal, G.D.; Ouyang, B.; Serrano, G.; Shill, H.A.; Goetz, C.; Stebbins, G.; Metman, L.V.; Driver-Dunckley, E.; Mehta, S.H.; Caviness, J.N.; et al. Comparison of Neuropathology in Parkinson’s Disease Subjects with and without Deep Brain Stimulation. Mov. Disord. 2017, 32, 274–277. [Google Scholar] [CrossRef]
  114. Fischer, D.L.; Manfredsson, F.P.; Kemp, C.J.; Cole-Strauss, A.; Lipton, J.W.; Duffy, M.F.; Polinski, N.K.; Steece-Collier, K.; Collier, T.J.; Gombash, S.E.; et al. Subthalamic Nucleus Deep Brain Stimulation Does Not Modify the Functional Deficits or Axonopathy Induced by Nigrostriatal α-Synuclein Overexpression. Sci. Rep. 2017, 7, 16356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Obeso, J.A.; Rodriguez-Oroz, M.C.; Rodriguez, M.; Macias, R.; Alvarez, L.; Guridi, J.; Vitek, J.; DeLong, M.R. Pathophysiologic Basis of Surgery for Parkinson’s Disease. Neurology 2000, 55, S7–S12. [Google Scholar] [PubMed]
  116. Maesawa, S.; Kaneoke, Y.; Kajita, Y.; Usui, N.; Misawa, N.; Nakayama, A.; Yoshida, J. Long-Term Stimulation of the Subthalamic Nucleus in Hemiparkinsonian Rats: Neuroprotection of Dopaminergic Neurons. J. Neurosurg. 2004, 100, 679–687. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Gratwicke, J.; Kahan, J.; Zrinzo, L.; Hariz, M.; Limousin, P.; Foltynie, T.; Jahanshahi, M. The Nucleus Basalis of Meynert: A New Target for Deep Brain Stimulation in Dementia? Neurosci. Biobehav. Rev. 2013, 37, 2676–2688. [Google Scholar] [CrossRef] [PubMed]
  118. Hotta, H.; Kagitani, F.; Kondo, M.; Uchida, S. Basal Forebrain Stimulation Induces NGF Secretion in Ipsilateral Parietal Cortex via Nicotinic Receptor Activation in Adult, but Not Aged Rats. Neurosci. Res. 2009, 63, 122–128. [Google Scholar] [CrossRef]
  119. Kilgard, M.P.; Merzenich, M.M. Cortical Map Reorganization Enabled by Nucleus Basalis Activity. Science 1998, 279, 1714–1718. [Google Scholar] [CrossRef]
  120. Leplus, A.; Lauritzen, I.; Melon, C.; Kerkerian-Le Goff, L.; Fontaine, D.; Checler, F. Chronic Fornix Deep Brain Stimulation in a Transgenic Alzheimer’s Rat Model Reduces Amyloid Burden, Inflammation, and Neuronal Loss. Brain Struct. Funct. 2019, 224, 363–372. [Google Scholar] [CrossRef]
  121. Li, Y.; Liu, L.; Barger, S.W.; Griffin, W.S.T. Interleukin-1 Mediates Pathological Effects of Microglia on Tau Phosphorylation and on Synaptophysin Synthesis in Cortical Neurons through a P38-MAPK Pathway. J. Neurosci. 2003, 23, 1605–1611. [Google Scholar] [CrossRef] [Green Version]
  122. Laxton, A.W.; Tang-Wai, D.F.; McAndrews, M.P.; Zumsteg, D.; Wennberg, R.; Keren, R.; Wherrett, J.; Naglie, G.; Hamani, C.; Smith, G.S.; et al. A phase I trial of deep brain stimulation of memory circuits in Alzheimer's disease. Ann. Neurol. 2010, 68, 521–534. [Google Scholar] [CrossRef]
  123. Lozano, A.M.; Fosdick, L.; Chakravarty, M.M.; Leoutsakos, J.M.; Munro, C.; Oh, E.; Drake, K.E.; Lyman, C.H.; Rosenberg, P.B.; Anderson, W.S.; et al. A Phase II Study of Fornix Deep Brain Stimulation in Mild Alzheimer’s Disease. J. Alzheimer’s Dis. 2016, 54, 777–787. [Google Scholar] [CrossRef] [Green Version]
  124. Hadar, R.; Bikovski, L.; Soto-Montenegro, M.L.; Schimke, J.; Maier, P.; Ewing, S.; Voget, M.; Wieske, F.; Götz, T.; Desco, M.; et al. Early Neuromodulation Prevents the Development of Brain and Behavioral Abnormalities in a Rodent Model of Schizophrenia. Mol. Psychiatry 2017, 23, 943–951. [Google Scholar] [CrossRef] [PubMed]
  125. Zeiss, C.J.; Shin, D.; Wyk, B.V.; Beck, A.P.; Zatz, N.; Sneiderman, C.A.; Kilicoglu, H. Menagerie: A Text-Mining Tool to Support Animal-Human Translation in Neurodegeneration Research. PLoS ONE 2019, 14, e0226176. [Google Scholar] [CrossRef] [PubMed]
  126. Konnova, E.A.; Swanberg, M. Animal Models of Parkinson’s Disease. In Parkinson’s Disease: Pathogenesis and Clinical Aspects; Codon Publications: Brisbane, Australia, 2018; pp. 83–106. [Google Scholar] [CrossRef]
  127. Anselmi, C.; Kowarsky, M.; Gasparini, F.; Caicci, F.; Ishizuka, K.J.; Palmeri, K.J.; Raveh, T.; Sinha, R.; Neff, N.; Quake, S.R.; et al. Two Distinct Evolutionary Conserved Neural Degeneration Pathways Characterized in a Colonial Chordate. Proc. Natl. Acad. Sci. USA 2022, 119, e2203032119. [Google Scholar] [CrossRef] [PubMed]
  128. Sekiya, M.; Wang, M.; Fujisaki, N.; Sakakibara, Y.; Quan, X.; Ehrlich, M.E.; De Jager, P.L.; Bennett, D.A.; Schadt, E.E.; Gandy, S.; et al. Integrated Biology Approach Reveals Molecular and Pathological Interactions among Alzheimer’s Aβ42, Tau, TREM2, and TYROBP in Drosophila Models. Genome Med. 2018, 10, 26. [Google Scholar] [CrossRef] [PubMed]
  129. Gomez-Tames, J.; Asai, A.; Hirata, A. Significant Group-Level Hotspots Found in Deep Brain Regions during Transcranial Direct Current Stimulation (TDCS): A Computational Analysis of Electric Fields. Clin. Neurophysiol. 2020, 131, 755–765. [Google Scholar] [CrossRef] [PubMed]
  130. Karimi, F.; Attarpour, A.; Amirfattahi, R.; Nezhad, A.Z. Computational Analysis of Non-Invasive Deep Brain Stimulation Based on Interfering Electric Fields. Phys. Med. Biol. 2019, 64, 235010. [Google Scholar] [CrossRef]
  131. Missey, F.; Rusina, E.; Acerbo, E.; Botzanowski, B.; Trébuchon, A.; Bartolomei, F.; Jirsa, V.; Carron, R.; Williamson, A. Orientation of Temporal Interference for Non-Invasive Deep Brain Stimulation in Epilepsy. Front. Neurosci. 2021, 15, 656. [Google Scholar] [CrossRef]
  132. Bocci, T.; Caleo, M.; Tognazzi, S.; Francini, N.; Briscese, L.; Maffei, L.; Rossi, S.; Priori, A.; Sartucci, F. Evidence for Metaplasticity in the Human Visual Cortex. J. Neural Transm. 2013, 121, 221–231. [Google Scholar] [CrossRef]
  133. Stagg, C.J.; Best, J.G.; Stephenson, M.C.; O’Shea, J.; Wylezinska, M.; Kineses, Z.T.; Morris, P.G.; Matthews, P.M.; Johansen-Berg, H. Polarity-Sensitive Modulation of Cortical Neurotransmitters by Transcranial Stimulation. J. Neurosci. 2009, 29, 5202–5206. [Google Scholar] [CrossRef] [Green Version]
  134. Bikson, M.; Dmochowski, J. What It Means to Go Deep with Non-Invasive Brain Stimulation. Clin. Neurophysiol. 2020, 131, 752–754. [Google Scholar] [CrossRef]
Figure 1. Schematic overview of neuroprotective effects of tDCS in animal models of Parkinson’s disease. SOD = superoxide dismutase; GSH-Px = glutathione peroxidase; BDNF = brain-derived neurotrophic factor; MDA = malonaldehyde; DA = dopamine; TH = tyrosine hydroxylase. Feng et al., 2020 [44]; Lee et al., 2019 [17]; Li et al., 2015 [19]; Lee et al., 2018 [45].
Figure 1. Schematic overview of neuroprotective effects of tDCS in animal models of Parkinson’s disease. SOD = superoxide dismutase; GSH-Px = glutathione peroxidase; BDNF = brain-derived neurotrophic factor; MDA = malonaldehyde; DA = dopamine; TH = tyrosine hydroxylase. Feng et al., 2020 [44]; Lee et al., 2019 [17]; Li et al., 2015 [19]; Lee et al., 2018 [45].
Ijms 23 13775 g001
Figure 2. Schematic overview of neuroprotective effects of NIBS in Alzheimer’s disease. tDCS = transcranial direct current stimulation; tACS = transcranial alternating current stimulation; rTMS = repetitive transcranial magnetic stimulation; BDNF = brain-derived neurotrophic factor; TrkB = Tropomyosin receptor kinase B; NGF = Nerve growth factor; NO = nitric oxide; OxS = oxidative stress. Choung et al., 2021 [70]; Velioglu et al., 2021 [71]; Tan et al., 2013 [73]; Chen et al., 2020 [74]; Marceglia et al., 2016 [83]; Luo et al., 2022 [88]; Khedr et al., 2019 [90].
Figure 2. Schematic overview of neuroprotective effects of NIBS in Alzheimer’s disease. tDCS = transcranial direct current stimulation; tACS = transcranial alternating current stimulation; rTMS = repetitive transcranial magnetic stimulation; BDNF = brain-derived neurotrophic factor; TrkB = Tropomyosin receptor kinase B; NGF = Nerve growth factor; NO = nitric oxide; OxS = oxidative stress. Choung et al., 2021 [70]; Velioglu et al., 2021 [71]; Tan et al., 2013 [73]; Chen et al., 2020 [74]; Marceglia et al., 2016 [83]; Luo et al., 2022 [88]; Khedr et al., 2019 [90].
Ijms 23 13775 g002
Figure 3. Schematic overview of comparisons in neuroprotective effects between NIBS and DBS. tDCS = transcranial direct current stimulation; DA = dopamine; SOD = superoxide dismutase; GSH-PX = glutathione peroxidase; TH = tyrosine hydroxylase; MPTP = 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine; MDA = malonaldehyde; LF-rTMS = low frequency repetitive transcranial magnetic stimulation; BDNF = Brain-Derived Neurotrophic Factor; NGF = nerve growth factor; NO = nitric oxide; Aβ = amyloid-β peptide; BBB = blood brain barrier; HF-rTMS = high frequency repetitive transcranial magnetic stimulation; TrkB = Tropomyosin receptor kinase B; SNpc = substantia nigra pars compacta; PAG = periaqueductal grey matter; Akt = Protein kinase B; STN = subthalamic nucleus. Feng et al., 2020 [44]; Lee et al., 2019 [17]; Li et al., 2015 [19]; Lee et al., 2018 [45]; Tan et al., 2013 [73]; Marceglia et al., 2016 [83]; Khedr et al., 2019 [90]; Luo et al., 2022 [88]; Chen et al., 2020 [74]; Choung et al., 2021 [70]; Velioglu et al., 2021 [71]; Wallace et al., 2007 [100]; Musacchio et al., 2017 [101]; Spieles-Engemann et al., 2011 [102]; Fischer et al., 2017 [103]; Rodrigues et al., 1998 [104]; Baldermann et al., 2018 [105]; Huang et al., 2019 [106]; Xia et al., 2017 [107]; Akwa et al., 2017 [108].
Figure 3. Schematic overview of comparisons in neuroprotective effects between NIBS and DBS. tDCS = transcranial direct current stimulation; DA = dopamine; SOD = superoxide dismutase; GSH-PX = glutathione peroxidase; TH = tyrosine hydroxylase; MPTP = 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine; MDA = malonaldehyde; LF-rTMS = low frequency repetitive transcranial magnetic stimulation; BDNF = Brain-Derived Neurotrophic Factor; NGF = nerve growth factor; NO = nitric oxide; Aβ = amyloid-β peptide; BBB = blood brain barrier; HF-rTMS = high frequency repetitive transcranial magnetic stimulation; TrkB = Tropomyosin receptor kinase B; SNpc = substantia nigra pars compacta; PAG = periaqueductal grey matter; Akt = Protein kinase B; STN = subthalamic nucleus. Feng et al., 2020 [44]; Lee et al., 2019 [17]; Li et al., 2015 [19]; Lee et al., 2018 [45]; Tan et al., 2013 [73]; Marceglia et al., 2016 [83]; Khedr et al., 2019 [90]; Luo et al., 2022 [88]; Chen et al., 2020 [74]; Choung et al., 2021 [70]; Velioglu et al., 2021 [71]; Wallace et al., 2007 [100]; Musacchio et al., 2017 [101]; Spieles-Engemann et al., 2011 [102]; Fischer et al., 2017 [103]; Rodrigues et al., 1998 [104]; Baldermann et al., 2018 [105]; Huang et al., 2019 [106]; Xia et al., 2017 [107]; Akwa et al., 2017 [108].
Ijms 23 13775 g003
Table 1. Studies assessing neuroprotective effects of tDCS in animal models of Parkinson’s disease.
Table 1. Studies assessing neuroprotective effects of tDCS in animal models of Parkinson’s disease.
StudySample/AnimalsPolarityConfigurationParametersBiological OutcomesBiological Results
Li et al., 2015 [19]36 C57Bl mice (n = 9 in control group; n = 9 in sham tDCS group; n = 9 in tDCS groups; n = 9 in drug group)Anodal/ShamAE: left frontal cortex;
R: between the shoulders
0.2 mA, 10 min/day, 21 consecutive days
AEA: 3.5 mm2
CD: 5.7 mA/cm2
DA, TH, SOD and GSH-PX activities, nonenzymatic MDA activitytDCS increased DA, SOD and GSH-Px; after MPTP induction, anodal tDCS increased TH and reduced MDA
Lee et al., 2018 [45]60 Male C57BL/6 mice (n = 15 in control group; n = 15 in anodal tDCS group; n = 15 in MPTP group; n = 15 in MPTP + tDCS group)Anodal/ShamAE: left motor cortex;
R: between the shoulders
0.1 mA, 30 min/day, 5 consecutive days
AEA: 3.1 mm2
CD: 3.2 mA/cm2
TH-positive cells; TH; α-synuclein protein; loss of dopaminergic neuron cells; ratio of LC3-II/LC3-I; p62; PI3K; mTOR; AMPK; ULKtDCS attenuated decrease of TH, p62, mTOR, PI3K, BDNF; attenuated increase of α-synuclein, LC3-II/LC3-I, AMPK and ULK
Lee et al., 2019 [17]Male C57BL/6 mice (number n.r.)Anodal/ShamAE: on motor cortex;
R: between the shoulders
0.1 mA, 30 min/day, 5 days/week, 1 week;
AEA: 3.1 mm2
CD: 3.2 mA/cm2
Expression of:
TH, mitophagy-related proteins; marker of degradation phase of autophagy; mitochondrial biogenesis-related proteins; mitochondrial fission and fusion -related proteins;
ATP concentration. Mitochondrial GDH activity
tDCS preserved neurons and fibers in substantia nigra and striatum;
attenuated mitochondrial GDH activity, ATP concentration; increased mitophagy-related and mitochondrial biogenesis proteins
Feng et al., 2020 [44]16 male Wistar (n = 8 in anodal group; n = 8 in sham group)Anodal/ShamAE: skull bregma;
R: anterior chest
300 μA, 20 min/day, 5 days/week, 4 weeks;
AEA: 37.9 mm2;
CD: 0.16 mA/cm2
Loss of dopaminergic nigrostriatal neurons and fiberstDCS preserved neurons in the substantia nigra, but not fibers in the striatum
tDCS = transcranial direct current stimulation; AE = active electrode; R = reference; AEA = active electrode area; CD = current density; DA = dopamine; TH = tyrosine hydroxylase; SOD = superoxide dismutase; GSH-PX = glutathione peroxidase; MDA = malonaldehyde; MPTP = 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine; LC3 = microtubule-associated protein 1 light chain 3; p62 = sequestosome 1/p62; PI3K = phosphoinositide 3-kinases; mTOR = mechanistic target of rapamycin; AMPK = AMP-activated protein kinase; ULK = unc-51-like kinase 1; ATP = adenosine triphosphate; GDH = glutamate dehydrogenase.
Table 2. Studies assessing neuroprotective effects of NIBS in animal models of Alzheimer’s Disease (AD).
Table 2. Studies assessing neuroprotective effects of NIBS in animal models of Alzheimer’s Disease (AD).
StudyNIBS
Method
Sample/AnimalsConfigurationParametersBiological OutcomesBiological Results
Tan et al., 2013 [73]rTMS (LF)84 mice (n = 21 in control group; n = 21 rTMS group; n = 21 in Aβ injection; n = 21 Aβ injection + rTMS)Whole brain stimulation400 pulses per session, 7 days/week, 2 weeks

LF-rTMS: 20 trains (20 pulses at 1 Hz, 10 s inter-interval)
Neuroplasticity-related proteins (BDNF, NGF and NMDA receptor) levelsLF-rTMS reversed NMDA receptor suppression, enhanced, BDNF and NGF levels
Marceglia et al., 2016 [83]tDCS (anodal/sham)7 AD patients (n = 7 tDCS; n = 7 sham)AE: bilateral temporo-parietal area;
R: right arm
1.5 mA, 15 min/day, 1 day

AEA: 25 cm2
CD: 0.06 mA/cm2
total NO levelstDCS increased NO levels
Khedr et al., 2019 [90]tDCS (anodal)46 AD patients
(n = 23 tDCS; n = 23 sham)
AE: bilateral temporo-parietal area;
R: left arm
2 mA, 20 min each side (5 min in between), 5 days/week, 2 weeks

AEA: 35 cm2
CD = 0.057 mA/cm2
AD brain damage biomarkers levels (TAU and Aβ 1-42)tDCS increased Aβ 1-42
Chen et al., 2020 [74]rTMS (HF)30 mice (n = 15 rTMS; n = 15 sham)Whole brain stimulation,600 pulses per session, 7 days/week, 2 weeks

HF-rTMS 20 trains (30 pulses at 5 Hz, 2 s inter-interval)
Synaptic plasticity-related proteins (PSD95), neurotrophic factors (BDNF, TrkB and AKT), autophagy marker proteins (p62 and LC3-II/LC3-I)HF-rTMS increased BDNF and TrkB levels, and enhanced hippocampal cellular autophagy
Choung et al., 2021 [70]rTMS
(HF/LF/sham)
24 mice (n = 8 HF-rTMS; n = 8 LF-rTMS; n = 8 sham)Whole brain stimulation1600 pulses per session, 5 days/week, 2 weeks

HF-rTMS: 40 trains (2 s duration at 20 Hz, 28 s inter-interval)

LF-rTMS: continuous stimulation (1 Hz).
BDNF, nestin and neuron protein levelsHF-rTMS increased BDNF, nestin and neuron expression levels in hippocampus and cortex, compared to sham
Velioglu et al., 2021 [71]rTMS (HF)15 subjectsLeft parietal cortex stimulation1640 pulses per session, 5 days/week, 2 weeks

HF-rTMS: 42 trains (2 s duration at 20 Hz, 28 s inter-interval)
BDNF and anti-oxidative stress proteins levelsHF-rTMS increased BDNF and anti-oxidative stress proteins levels
Luo et al., 2022 [88]tDCS (anodal/sham)33 AD model mice (n = 11 tDCS; n = 11 not treated; n = 11 sham)AE: frontal cortex;
R: thorax
150 µA, 30 min/day, 5 days/week, 2 weeks

AEA:nr
CD: nr
Aβ plaques density in the hippocampus and frontal cortex,
NVU integrity
tDCS reduced Aβ plaques density and increased
BBB integrity-related proteins
LF-rTMS = low frequency repetitive transcranial magnetic stimulation; Aβ = amyloid-β peptide; BDNF = Brain-Derived Neurotrophic Factor; NGF = Nerve growth factor; NMDA = N-methyl-D-aspartate receptor; tDCS = transcranial direct current stimulation; AD = Alzheimer’s disease; AE = active electrode; R = reference; AEA = active electrode area; CD = current density; NO = nitric oxide; TAU = tubulin associated unit; HF-rTMS = high frequency repetitive transcranial magnetic stimulation; PSD95 = Postsynaptic density protein 95; TrkB = Tropomyosin receptor kinase B; AKT = protein kinase B; LC3 = microtubule-associated protein 1 light chain 3; p62 = sequestosome 1/p62; NVU = neurovascular unit; BBB = blood brain barrier.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Guidetti, M.; Bertini, A.; Pirone, F.; Sala, G.; Signorelli, P.; Ferrarese, C.; Priori, A.; Bocci, T. Neuroprotection and Non-Invasive Brain Stimulation: Facts or Fiction? Int. J. Mol. Sci. 2022, 23, 13775. https://doi.org/10.3390/ijms232213775

AMA Style

Guidetti M, Bertini A, Pirone F, Sala G, Signorelli P, Ferrarese C, Priori A, Bocci T. Neuroprotection and Non-Invasive Brain Stimulation: Facts or Fiction? International Journal of Molecular Sciences. 2022; 23(22):13775. https://doi.org/10.3390/ijms232213775

Chicago/Turabian Style

Guidetti, Matteo, Alessandro Bertini, Francesco Pirone, Gessica Sala, Paola Signorelli, Carlo Ferrarese, Alberto Priori, and Tommaso Bocci. 2022. "Neuroprotection and Non-Invasive Brain Stimulation: Facts or Fiction?" International Journal of Molecular Sciences 23, no. 22: 13775. https://doi.org/10.3390/ijms232213775

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