Skip to main content

Electrical stimulation methods and protocols for the treatment of traumatic brain injury: a critical review of preclinical research

Abstract

Background

Traumatic brain injury (TBI) is a leading cause of disabilities resulting from cognitive and neurological deficits, as well as psychological disorders. Only recently, preclinical research on electrical stimulation methods as a potential treatment of TBI sequelae has gained more traction. However, the underlying mechanisms of the anticipated improvements induced by these methods are still not fully understood. It remains unclear in which stage after TBI they are best applied to optimize the therapeutic outcome, preferably with persisting effects. Studies with animal models address these questions and investigate beneficial long- and short-term changes mediated by these novel modalities.

Methods

In this review, we present the state-of-the-art in preclinical research on electrical stimulation methods used to treat TBI sequelae. We analyze publications on the most commonly used electrical stimulation methods, namely transcranial magnetic stimulation (TMS), transcranial direct current stimulation (tDCS), deep brain stimulation (DBS) and vagus nerve stimulation (VNS), that aim to treat disabilities caused by TBI. We discuss applied stimulation parameters, such as the amplitude, frequency, and length of stimulation, as well as stimulation time frames, specifically the onset of stimulation, how often stimulation sessions were repeated and the total length of the treatment. These parameters are then analyzed in the context of injury severity, the disability under investigation and the stimulated location, and the resulting therapeutic effects are compared. We provide a comprehensive and critical review and discuss directions for future research.

Results and conclusion

We find that the parameters used in studies on each of these stimulation methods vary widely, making it difficult to draw direct comparisons between stimulation protocols and therapeutic outcome. Persisting beneficial effects and adverse consequences of electrical simulation are rarely investigated, leaving many questions about their suitability for clinical applications. Nevertheless, we conclude that the stimulation methods discussed here show promising results that could be further supported by additional research in this field.

Background

Most recent epidemiological surveillance reports indicate around 223,000 traumatic brain injury related hospitalizations in 2019 and more than 64,000 TBI-related deaths in 2020 in the USA alone [1]. Recent analysis of data from the European Union in 2017 shows a much higher number of TBI-related hospitalizations, although there are less TBI-related deaths [2]. Despite substantial differences among countries, TBI remains a leading cause of mortality and morbidity, particularly amongst the younger population. Decades-long accumulation of clinical and experimental data has set the path to considerable achievements in the clinical management of TBI, which brought a remarkable and gradual reduction in mortality due to head injury [3, 4]. Nevertheless, neurological deficits, cognitive and motor impairments, psychiatric disorders or other morbidities remain among the major sequelae of TBI [5]. Whilst these disabilities render TBI survivors dependent on assistance for daily activities, they also cause severe psychological and economic burden on families due to lifelong patient care.

A modest list of major pathophysiological changes after TBI includes dysregulated cerebral blood flow [6] and impaired cerebral oxygenation leading to ischemic insult [7], glutamate excitotoxicity [8, 9], blood brain barrier breakdown [10], cerebral edema [11, 12], oxidative and nitrosative stress [13, 14], cerebral inflammation [15, 16], hypo- and hyper perfusion [17], mitochondrial dysfunction [18], hemorrhage [19] and hyperemia [20]. The cascade of these pathophysiological changes starts within minutes to hours and days following the primary injury, and may directly or indirectly induce secondary damage to brain tissue, resulting in impaired connectivity and a delayed loss of neuronal cells. Moreover, chronic microglial activation and axonal damage may persist over much longer periods, leading to connectivity loss even years after trauma [21]. Based on the order of appearance of those pathologies, the post-TBI period can be roughly divided into the acute phase lasting minutes to hours after trauma, the subacute phase that lasts several days and is connected to the beginning of the secondary injury, and the chronic phase covering the weeks, months or even years following TBI [22, 23]. Decades of immense clinical and preclinical research were dedicated to deciphering the mechanisms of secondary damage and cell loss. Nevertheless, the continuously increasing knowledge in this field has not yet yielded the desired clinical applications for targeted pharmacological therapies to prevent or attenuate these mechanisms and stop further progression of tissue damage.

Neuromodulation by means of electrical and magnetic stimulation has been used to promote neuroplasticity and connectivity. Although the limited capabilities of nervous tissue to self-repair hinders complete regeneration of damaged brain tissue, processes involved in neuroplasticity can at least partially restore neuronal connectivity. Promising results observed in preclinical and clinical studies with electrical stimulation provide a good basis for the exploitation of neuroplasticity for functional restoration to alleviate trauma-induced disabilities [24, 25]. Hypo- or hyper-excitability, for instance, provide suitable targets for neuromodulatory interventions such as transcranial magnetic stimulation and deep brain stimulation [26, 27]. Supportive treatment of post-traumatic depression using electrical stimulation has also been subject to an immense interest [28, 29]. Preclinical studies, however, which are required to corroborate findings on underlying mechanisms of electrical stimulation and reveal neurobiological correlates of these stimulation techniques, are disproportionately sparse and appear to have attracted increased interest only over the last decade.

In the first part of this article, we give an overview on what is known about the effects of stimulation on neuronal cells and the state-of-the-art of the most commonly used electrical stimulation methods for therapeutic applications. In the second part we present a critical review of the available literature on preclinical studies using electrical stimulation in animal models of traumatic brain injury. The aims are (1) to assess the efficacy of these stimulation methods as post-TBI treatments in preclinical research across several selected studies, (2) to critically compare stimulation protocols as well as treatment time after traumatic insult and (3) to infer the translational value of the reported outcomes for clinical applications.

State-of-the-art

Effects of electrical stimulation on neurons

The excitability of neuronal cells facilitates modulation of their firing activity using external stimulation to enhance or suppress endogenous activity [30]. This modulation can be utilized for therapeutic or diagnostic purposes in several neurological diseases or injuries to nervous tissue [31,32,33,34]. To better understand the advantages of therapeutic electrical stimulation following TBI, it is necessary to gain extensive insights into how and to which extent stimulation influences neuronal physiology and morphology.

Artificial electrical stimulation may change the electrical potential of the surrounding extracellular region through anodic as well as cathodic protocols [35,36,37]. In cathodic stimulation, a negative current pulse is delivered to the extracellular area, which in turn depolarizes the cellular membrane with the aim to elicit an action potential. Anodic stimulation instead hyperpolarizes the region near the site of interest and thus decreases the membrane potential [38]. This results in a flux of positive ions towards the stimulation site from surrounding areas, which leads to a depolarization of the cell membrane further away from the site of stimulation, possibly triggering an action potential at the nearest Ranvier node [39, 40].

The effect of electrical stimulation on the brain depends on the inherent characteristics of the tissue. At the cellular level, it is easier to excite an axon than a soma, and myelinated axons are the most excitable part of the cell [41, 42]. Induced voltages differ between nodes and internodes due to the drastic differences in voltage-gated ion channel density [43, 44]. Activated axons progress the signal antidromically to the soma [45, 46] and orthodromically to the synaptic terminals [47, 48]. Bending, branching and significant changes in the diameter of an axon determine the effective site and threshold of the stimulation [43]. Generally, it is easier to elicit action potentials with negative currents in almost all cell compartments, except for some types of dendrites that are more prone to stimulation with positive currents [43, 49, 50].

Long-term potentiation (LTP), long-term depression (LTD) and plasticity

Electrical stimulation deeply influences brain electrophysiology through modulation of neuronal signaling not only in the short-term, but also in facilitating or attenuating long-term modifications on a cellular level [51, 52]. Activity-dependent synaptic plasticity can either strengthen or weaken the development of synapses [53, 54], which is crucial for post-traumatic regeneration and recuperation of high-level cognitive abilities like learning and memory formation, loss of which is a typical outcome of TBI [55, 56]. Long-term potentiation (LTP) and long-term depression (LTD) are highly complex and pivotal processes of synaptic plasticity, which may be heavily modified as a consequence of TBI, possibly leading to severe cognitive impairments [56].

LTP is a form of synaptic enhancement resulting in a lasting facilitation of signal transduction. Classically, LTP is elicited through brief high frequency stimulation [57], although it may also be triggered successfully with theta-burst stimulation protocols [58] or chemical compounds [59]. Initiation of LTP requires the activation of postsynaptic N-methyl-d-aspartate (NMDA) receptors with glutamate. Subsequent rapid increase of calcium concentration within the cell initiates multiple metabolic cascades and the modulation of gene transcription, resulting in long-term changes to receptor expression, synaptic vesicle transport and other cytoskeletal interactions [55, 60]. LTD is a process analogous to LTP, but leads to reduction of synaptic efficacy. It is usually induced by low frequency stimulation, leading from low to moderate influx of calcium into the neuron mostly through voltage-gated calcium channels and, to a lesser extent, through the activation of NMDA receptors [60, 61].

During a head trauma, mechanical forces applied to nervous tissue disturb ionic fluxes and the concomitant depolarization [62]. This leads to excessive glutamate release from presynaptic axon terminals in the acute phase of the injury that may result in neuronal hyperexcitability and changes in synaptic plasticity. In general, TBI attenuates synaptic LTP responses, while its effect on LTD may vary [56]. LTP deficits and overall increased neuronal excitability were observed soon after injury in in vivo and ex vivo TBI models [63, 64], while the ability to induce LTD was left unchanged [64]. In a controlled cortical impact model in rats, LTD was enhanced as long as 2 days after the initial injury [65]. Considering all of the above, pertinent electrical stimulation protocols to effectively modulate LTP and LTD could be advantageous in the recovery of physiological neuroplasticity mechanisms and the recuperation of impeded motor and cognitive functions following TBI.

Spike timing-dependent plasticity (STDP)

Timing of the activation of presynaptic and postsynaptic cells plays a pivotal role in synaptic plasticity [66, 67]. Spike timing-dependent plasticity (STDP) is considered a biologically plausible model for synaptic modifications occurring in vivo [68, 69] and its occurrence has been reported in several brain regions, such as the corticostriatal pathway [70,71,72], the barrel cortex [73, 74] and the visual cortex [75, 76]. It is determined by the temporal order of action potential initiations and the narrow time between subsequent action potentials. In general, activation of the presynaptic cell immediately before activation of the postsynaptic cell leads to timing-dependent LTP, while activating the presynaptic neuron immediately after the postsynaptic cell elicits timing-dependent LTD [68, 69]. The time window between these activations needs to be in the order of milliseconds, is specific for each synapse and depends on receptor kinetics, current densities and the release of retrograde messengers such as endocannabinoids [69]. Spontaneous spiking as well as changes in the spike frequency can further modulate the strength of plasticity, e.g. higher frequency stimulation has been described to increase the effect of timing-dependent LTP [69]. STDP was observed in both excitatory and inhibitory neurons and could be further modified by cholinergic, dopaminergic and adrenergic signaling [68], enabling prospective pharmacological modulation. It offers an alternative to frequency-dependent stimulation in clinical settings and has already been deployed in human studies to successfully modulate the force of the long-latency stretch reflex in healthy volunteers [77], while overall lower limb motor output was improved in patients with spinal cord injury [78].

Electrical stimulation methods

The most prevalent electrical stimulation methods used in post-TBI treatment studies, which are in the focus of this review, are transcranial magnetic stimulation (TMS), transcranial direct current stimulation (tDCS), deep brain stimulation (DBS) and vagus nerve stimulation (VNS) [25, 79]. TMS and tDCS are amongst the most commonly used non-invasive brain stimulation techniques [80, 81]. They are effective in the treatment of a wide variety of neurologic impairments, but their efficiency and precision is limited by the distance of the stimulator to the target region. Invasive stimulation methods, such as DBS or VNS, may achieve higher precision and efficiency by bringing the stimulation electrodes closer to the desired area. A schematic overview of these four stimulation methods and their preclinical usage is depicted in Fig. 1.

Fig. 1
figure 1

Simplified overview on preclinical applications of the four stimulation methods in the focus of this review: a Transcranial magnetic stimulation (TMS) uses magnetic fields to stimulate neurons in the brain non-invasively. b Transcranial direct current stimulation (tDCS) delivers low intensity electrical currents to the brain via scalp electrodes in order to modulate neural activity. c Deep brain stimulation (DBS) involves the implantation of a device that delivers electrical impulses to specific areas of the brain. d Vagus nerve stimulation (VNS) uses cuff electrodes to deliver electrical stimulation to the vagus nerve. Figure created with BioRender.com

Transcranial magnetic stimulation

TMS is a non-invasive method that utilizes magnetic fields to inhibit or enhance the electrical activity of brain tissue with the aim to improve various neurologic disabilities [82, 83]. This technique utilizes the physical principle of electromagnetic induction by running a high alternating current through a magnetic coil positioned tangentially to the skull of a subject, leading to the formation of a magnetic field that is able to penetrate the skull. When stimulation is applied in the form of pulses, the rapid changes in the magnetic field create electrical currents in the brain, which in turn leads to excitation or inhibition of electrical activity, depending on the frequency of stimulation [84].

The main limitation of this method is that the electromagnetic field created by the coil rapidly decreases in strength with increasing distance. Thus, TMS is mainly used to stimulate cortical areas near the surface of the brain; however, functionally connected regions deeper in the brain can be stimulated indirectly through projecting axons [24]. The depth that the magnetic field penetrates into the brain as well as the size of the stimulated area can be adapted to specific requirements by selecting different coil types with various geometries, materials and designs. Circular coils, for example, can be used to uniformly stimulate a larger volume of neuronal tissue, resulting in greater penetration depth. Figure-of-eight shaped designs, comprising two circular coils positioned next to each other, allow for more selective stimulation at the cost of penetration depth [85]. The area where the two electromagnetic fields produced by this arrangement overlap is characterized by an increased current density compared to the surrounding regions [84]. TMS can be applied in a wide variety of different protocols, most commonly in the form of repetitive pulses, which is usually referred to as repetitive TMS (rTMS) [86].

The therapeutic potential of rTMS is widely recognized, particularly in the field of psychiatry, and it is applied as a treatment option for depression [87, 88] and obsessive–compulsive disorder [89, 90]. Its efficacy was further tested as a treatment for a number of different neurological conditions, such as neuropathic pain [91, 92], epilepsy [93], stroke [94], multiple sclerosis [95] and post-traumatic stress disorder [96], as well as Parkinsonian movement disorders [97, 98].

Transcranial direct current stimulation

In contrast to other stimulation methods that employ pulsed protocols for neurostimulation, tDCS uses direct current to influence the cell membranes of neurons in the desired cortical area [99, 100]. A current of several milliamperes is applied via a pad electrode, called the active electrode, attached to the pericranium near the area of interest, which leads to changes in cortical excitability and neuronal activity [101, 102]. A second, larger reference electrode is usually placed further away from the stimulation site. During anodal tDCS, a positive current is applied between the two electrodes, leading to a hyperpolarization of the area near the active electrode, whilst cathodal tDCS depolarizes the tissue with the use of negative currents. The resulting excitation or inhibition of neurons may lead to neuromodulation in affected areas [99, 103]. The current density is crucial for the efficacy and propagation depth of the stimulus [104].

This method is painless, non-invasive and used as a treatment for depression and a variety of cognitive dysfunctions [105, 106]. However, lack of precision is a limiting factor in cases where targeted neurostimulation would be necessary, such as post-traumatic tremor [107].

Deep brain stimulation

DBS is an invasive approach that requires the implantation of a stimulation electrode directly into the targeted brain area [108, 109]. The stimulation setup comprises an implanted stimulation electrode and a connected subcutaneous wire that forwards signals from an external stimulating device. Stimulation electrodes are often implanted bilaterally and commonly have multiple metal contacts, which can be used both as anodes and as cathodes [110]. In bipolar configurations, an electrical field is generated between two adjacent contacts, allowing for a concentrated electric field and thus a higher precision [110]. The optimal electrode position is usually determined beforehand with the help of neuroimaging via computed tomography (CT) or magnetic resonance imaging (MRI), which can also be used to guide the surgeon during implantation. Throughout the procedure, electrical activity is continuously measured to ensure correct electrode placement. Afterwards, the efficacy of the implanted device is verified by applying initial stimulation pulses [111].

This method is approved for the symptomatic treatment of Parkinson’s disease, essential tremor, obsessive compulsive disorder and some cases of severe epilepsy in humans [112, 113]. Thanks to its versatility and high spatial resolution, DBS has potential use in the treatment of higher-order cognitive dysfunction and disorders of consciousness in patients with TBI [114].

Vagus nerve stimulation

VNS is an indirect brain stimulation method that excites the afferents of the vagus nerve to modulate activity of the central nervous system. While vagal afferents provide sensory information to the brain stem from multiple internal organs, efferents mediate the parasympathetic control of various bodily functions. Thus, VNS results in a wide range of different effects caused by the stimulation of medulla and brainstem including the modulation of neurotransmitters: notably epinephrine, serotonin and gamma-aminobutyric acid [115]. Other potential modes of action include changes in blood flow in several brain regions [116,117,118], upregulation of neurotrophin production [119], reduction of damage to the blood brain barrier [120,121,122] and anti-inflammatory effects [123, 124]. VNS systems are approved for treatment of drug-resistant epilepsy [125] and severe, recurrent unipolar and bipolar depression [126], both of which are common disorders developing as a consequence of TBI [127,128,129].

Most commonly, VNS is used as an invasive modality, employing helical cuff electrodes in monopolar, bipolar or tripolar configurations. These electrodes are usually wrapped around the left cervical vagus nerve [130] to indirectly stimulate distant brain regions. Stimulation of the right vagus nerve might lead to severe bradycardia and is therefore generally avoided [130]. Monopolar electrodes are comparatively cheap, but require an additional ground electrode. Bipolar configurations allow the induced current to flow between two electrodes, enabling a much greater control of the current path. Tripolar electrodes are more expensive, but have the advantage of preventing leakage currents to the surrounding tissue since the stimulating electrode is positioned between two common counter-electrodes.

Stimulation waveforms and protocols

The selection of suitable protocols is an important factor for efficacious stimulation, but also for the prevention of damage to the stimulating electrodes and the surrounding tissue [35]. This is particularly relevant for invasive approaches, such as DBS and VNS, where implanted electrodes need to last for longer periods of time and are in direct contact with neural tissue [131]. Unwanted electrochemical reactions at the electrode-tissue interface include corrosion and oxygen reduction reactions, which can be minimized by selecting appropriate stimulation protocols and waveforms [132, 133]. While monophasic pulses are more efficacious for stimulation than biphasic pulses, they potentially result in greater tissue damage, since all injected charge creates electrochemical reaction products and result in greater negative overpotentials over time [35, 134]. Biphasic stimulation, on the other hand, has the potential to reverse electrochemical processes at the electrode-tissue interface, but may also reverse some of the desired effects necessary for efficacious charge induction. Introducing a short interphase delay reduces the suppressing effect of the reversal phase, as long as the delay is short enough to prevent excessive accumulation of electrochemical reaction products [35].

Another important part of the stimulation protocol is the timing of the treatment application after injury, which depends on the selected treatment modality, the severity of the trauma and the goal of the treatment [135]. The onset of stimulation in preclinical studies varies from immediately to several weeks after trauma [25]. In clinical settings, these techniques are usually applied at later stages as a support to traditional rehabilitation methods for treating disabilities that persist after TBI [114, 136].

Additional stimulation methods

In addition to the methods mentioned above, there are several other promising electrical stimulation modalities that may be effective in the treatment of TBI sequelae. Electrical cortical stimulation, an invasive method where electrodes are implanted near the cortical surface, can be used to modulate brain plasticity to treat sensorimotor and cognitive deficits in rats [137]. Similarly, epidural electrical stimulation utilizes pulsed stimulation protocols applied to electrodes implanted in the epidural or subdural space to enhance motor recovery and brain activity [138,139,140]. Promising non-invasive TBI treatment methods include electroconvulsive therapy, which finds use as the treatment for mood disorders such as depression [141], but has not yet been investigated in preclinical TBI models.

Temporal interference stimulation is another novel treatment modality that can be used to stimulate deep brain regions non-invasively, exploiting a well-known acoustic phenomenon [142]. By applying two sinusoidal stimuli in the kilohertz-range with slightly differing frequencies through electrode pairs placed on the head of a patient, interference patterns can be generated inside the brain [143]. The effect of stimuli in the kilohertz range on the underlying tissue is only small due to the filtering properties of cellular membranes [144, 145], and the amplitude of the individual signals is comparably low. Constructive interference of these two signals in the target area leads to an electric field oscillating with an envelope frequency equal to the difference between the two individual signal frequencies. This method has successfully been applied to mouse motor cortex, leading to the elicitation of movements [146].

It is also possible to implant passive components in the brain that convert an external impulse from a source outside the skull into an electrical stimulus. An example for this would be photocapacitors [147,148,149], which charge up when they are irradiated by light pulses, creating an electric field at their surface, leading to the depolarization of adjacent neural cells. These photocapacitive devices can also be used in combination with temporal interference stimulation protocols [150].

Systematic literature review

To gain further insight into the methods and protocols used for TBI therapy in preclinical studies, an extensive systematic literature search was conducted. The articles included in this survey were found in PubMed and Web of Science. The scientific integrity of the review was ensured by closely following the PRISMA 2020 guidelines [151]. A flow diagram detailing the literature assessment process is given in Fig. 2.

Fig. 2
figure 2

PRISMA 2020 flow diagram depicting the selection process of the studies for this review [151] (TBI traumatic brain injury, ES electrical stimulation, TMS transcranial magnetic stimulation, tDCS transcranial direct current stimulation, DBS deep brain stimulation, VNS vagus nerve stimulation)

Search terms: literature identification

To cover the most commonly used variations that describe the stimulation methods selected for this review as well as TBI, the search query consisted of the following MeSH terms:

(“transcranial magnetic stimulation” OR “transcranial direct current stimulation” OR “deep brain stimulation” OR “vagus nerve stimulation” OR "vagal nerve stimulation") AND (“traumatic brain injury” OR “tbi” OR “concussion”)

The search was conducted in the PubMed and Web of Science databases. To obtain as many relevant records as possible, the query was searched in all fields of the respective databases, which includes titles, abstracts and keywords of publications, among other information. The list of search results was last updated on the 8th of September 2022 and the search yielded 358 results in PubMed and 524 in Web of Science, amounting to a total of 583 different records after removing duplicates. The results were sorted by publication date from oldest to most recent and the titles, authors and publication years of these records were exported from the respective databases and collected in a Microsoft Excel spreadsheet to organize further screening.

Inclusion criteria: literature screening

The abstracts of all 583 individual search results were screened by one investigator for five different criteria of interest to this review. This was done manually without the use of any advanced automation tools except for a simple text search function. First, the abstract needed to mention TBI as the underlying cause of the disability under investigation. Next, an electrical stimulation method had to be utilized in the study and third it had to be used for a therapeutic purpose or as a treatment, as opposed to a diagnostic application. The record also needed to consist of original research, which excluded other review articles and excerpts from larger studies, such as meeting abstracts and conference papers. Finally, only preclinical studies were included, where an animal model was utilized to investigate certain parameters of interest.

These five criteria were assessed in the order described above, and when an article did not contain that criterion, it was immediately excluded from the review. A total of 543 records were excluded, 93 of which did not investigate TBI sequelae, 60 were not about electrical stimulation, 97 used these methods for an application other than therapy, 162 were not original research and 131 of the remaining articles were not preclinical studies. This abstract screening resulted in 40 articles for the following full-text assessment step.

Eligibility: full-text assessment

Out of the 40 articles selected for full-text assessment, another six were excluded. Four of the excluded articles used electrical stimulation not for the therapy but for the assessment of stimulation effects on healthy animals. One study was not original research, which was not immediately apparent in the abstract, and another article did not utilize electrical stimulation altogether. Ultimately, literature screening led to a total of 34 articles that were reviewed in this study. Eight separate studies used TMS and VNS respectively, seven employed tDCS, and eleven utilized DBS for the treatment of TBI sequelae.

Results

During full-text assessment, multiple parameters were collected from the 34 selected articles for further analysis and comparison. The first two columns list general information about the respective study, such as its main focus and the impairment under investigation. The next column describes the animal model used in each study, which includes the number and type of animals, the applied TBI model, and if animals were anesthetized during stimulation. After that, the technical aspects of the applied stimulation are summarized, such as the stimulation protocol that was used, the time frame of the stimulation, and the location that was stimulated. The last set of parameters focuses on the assessment of the study results, namely the tests that were conducted with the animals, the parameters that were studied, if they observed any long-term effects of stimulation, and a short summary of the main findings of the paper. All this information was collected in four individual tables, one each for TMS, tDCS, DBS and VNS, which are displayed below (Tables 1, 2, 3, 4).

Table 1 Overview of preclinical transcranial magnetic stimulation (TMS) studies
Table 2 Overview of preclinical transcranial direct current stimulation (tDCS) studies
Table 3 Overview of preclinical deep brain stimulation (DBS) studies
Table 4 Overview of preclinical vagus nerve stimulation (VNS) studies

Transcranial magnetic stimulation

Most of the included TMS studies listed in Table 1 investigated the loss of motor functions after TBI [138, 152,153,154,155], while some also used it as a potential treatment for detrimental changes in brain metabolism [152, 156], behavioral impairments [157], and to prevent cell death [152, 156]. A recent study also investigated the mechanisms of rTMS treatment without considering any specific disability [158]. Animals were usually immobilized and awake during stimulation, except for two studies, where TMS was applied during temporary anesthesia using volatile anesthetics [153, 154]. In four of the studies stimulation was done at the ipsilateral side [138, 152, 157, 158], in one at the contralateral side [153], and in two at the medial alignment to the injury site [154, 156]. Stimulus intensities are rarely given as absolute values, such as a magnetic field strength, but as a percentage of the maximum output of the stimulator [138, 153, 156, 157] or of the experimentally determined resting motor threshold of the animal [152, 158], while two studies do not specify the intensity of the stimulation [154, 155]. Many protocols employed rTMS in the form of pulse trains at frequencies ranging from 2 to 40 Hz, some of them having applied the stimulus for 9 to 20 min [138, 153, 155,156,157,158], while others stimulated for 3 min or less [152, 154]. In half of the studies TMS treatment was started 1 day after injury [138, 154, 156, 157], and the other half started stimulation several days later [152, 153, 155, 158]. Stimulation sessions were usually administered daily and continued for 1 week or longer. The target of TMS was often a nonspecific area of the cortex, apart from one study where the primary sensory region on the non-injured side of the brain was stimulated in pediatric animals [153] and another study that specifically targeted subcortical areas [155]. Persisting effects of TMS were rarely investigated, but one research group claims to have found a long-lasting increase of excitability in the non-injured cortex [153], while another found functional improvements lasting for up to 6 weeks after stimulation when TMS was combined with environmental enrichment [154]. Three studies observed a neuroprotective effect and the prevention of cell death [152, 155, 156], while two each determined that TMS could be an effective treatment to improve motor function [138, 155], induce neural plasticity [153, 158], or help with the recovery of brain activity [138, 152]. It was also shown that TMS led to histologic improvements after TBI, meaning that the expression levels of relevant proteins changed towards a positive outcome [155, 157, 158]. Individual studies determined that TMS could decrease hyperactivity [153], improve cell metabolism and at the same time induce cell proliferation and neurogenesis [156], help with the recovery from behavioral impairments [157], improve cortical excitability [154], or enhance cognitive function [155]. Only one study did not observe any improvements in motor function after applying TMS [152].

Transcranial direct current stimulation

The studies shown in Table 2 used tDCS mainly to assess improvements in motor function, excitability and cognitive impairments [159,160,161,162,163], but also its effects on cerebral blood flow (CBF) and tissue oxygenation after TBI [161, 164]. Only one study examined tDCS as a treatment for psychiatric-like symptoms such as impulsivity and attention [165]. Animals were anesthetized during tDCS in four of the seven studies [159, 160, 163, 165] and stimulation was applied for 10–30 min in all studies. In four studies, sessions were repeated for several days and lasted up to 4 weeks [159, 161, 162, 165], while three studies applied the stimulation only once in either the acute [160], subacute [163] or chronic phase [164] after TBI respectively. In six of the seven studies [159,160,161,162,163,164] anodal tDCS with an amplitude between 0.1 and 0.2 mA was applied. Nevertheless, the surface area of the employed electrodes varied considerably, resulting in widely different current densities between 0.255 and 2.82 mA/cm2, which is a critical factor for effective stimulation [104]. The anode was usually placed near the lesion or motor cortex, and the cathode at the thorax or trunk of the animal. Only one study [165] employed cathodal instead of anodal tDCS with a higher amplitude of 0.8 mA, resulting in a current density of 0.708 mA/cm2, whereby the cathode was placed near the bregma and the anode between the scapulae. One group observed a persisting increase in local cortical CBF in response to tDCS in TBI and control animals, as well as improved motor and cognitive outcome 1 week after the end of the stimulation in one of the stimulation groups [161]. However, all other studies in this scope that investigated long-term changes after stimulation [159, 160, 165] found that the beneficial effects of the treatment were no longer apparent after longer observation periods, over which non-treated animals reached a similar level of recovery.

Deep brain stimulation

With the possibility to target small and specific areas as well as deeper regions of the brain, DBS can be used to treat a wide variety of different impairments, such as the loss of cognitive [166,167,168,169,170] and motor function [171], as well as bladder dysfunction [172,173,174] and disorders of consciousness [175, 176]. Whilst the stimulation protocols differ greatly with respect to the targeted region and TBI sequelae, as shown in Table 3, the analyzed studies invariably reported positive results. Animals were generally kept awake during DBS, unless the stimulation was applied simultaneously with cystometric assessments [172,173,174]. Some studies utilized a current-controlled approach with amplitudes ranging from 20 to 200 µA [166,167,168,169, 175] or at 80% of the individual resting motor threshold [170, 171], while others applied voltages between 1 and 4 V [172,173,174, 176]. A stimulation frequency of 50 Hz seemed to be effective in the treatment of bladder dysfunction [172,173,174], while lower frequencies were used to treat motor [171] and cognitive deficits [166,167,168, 170], and higher frequencies of up to 200 Hz can be employed to increase arousal [175, 176]. Task-matched stimulation at 130 Hz for 5 s after each successful trial in a spatial learning test was also used to treat cognitive impairments after TBI [169]. In two studies, stimulation was applied directly before cognitive tests [166, 168], while, in the treatment of bladder dysfunction, stimulation was only triggered during cystometry when the measured bladder pressure exceeds a certain threshold [172,173,174]. Three studies applied stimulation over 12 daylight hours over several consecutive days to improve spatial memory [167, 170, 171], and two others investigating the potential of DBS to increase arousal started their continuous stimulation protocols directly after TBI over the course of 2 h to 1 day [175, 176]. The targeted brain area and stimulation onset highly depend on the treatment application in question, since DBS can be used to stimulate relatively small brain regions—compared to other stimulation methods—without affecting the surrounding tissue. Long-lasting effects of DBS were only reported in [169], where researchers observed improved recovery of spatial memory 10 days after cessation of stimulation compared to untreated animals; meanwhile, other studies reported that they did not find persisting effects on hippocampal theta power after stimulation was terminated [166, 168].

Vagus nerve stimulation

VNS has been used in the preclinical studies listed in Table 4 to improve motor and cognitive impairments [177,178,179] as well as disorders of consciousness [180, 181] after TBI, but also in the treatment of cerebral edema [182, 183] and to prevent cell death [184]. Animals were usually awake during VNS, except in two studies where researchers intentionally anesthetized animals to investigate the effect of VNS on disorders of consciousness [180, 181]. One study does not state clearly whether animals were anesthetized during the VNS or not [183]. Four studies applied stimuli at an amplitude of 0.5 mA and a frequency of 20 Hz [177, 178, 182, 184], while three other studies used currents between 0.8 and 1 mA with a frequency of 30 Hz [179,180,181, 183], all of which chose to stimulate the left vagus nerve at the cervical level. Stimulation was often applied for 30 s in 30 min intervals over a period of up to 2 weeks, starting within 2 [177, 182] or 24 h after injury [178, 184], while two studies applied the stimulation only once, directly after induction of TBI [180, 181]. In one of the studies, stimulation was applied for 500 ms within 45 ms after each successful trial in a pull performance task, with the aim to improve motor function [179]. Only in one study stimulation was applied to the right vagus nerve at a frequency of 5 Hz with 5 ms pulses and an amplitude of 10 V, once for 20 min, in an effort to alleviate brain edema [183]. Most of the studies in this scope did not investigate any possible persisting effects, since VNS is mostly used as a continuous treatment after injury. The study conducted by Pruitt et al. measured persisting effects 1 week after the completion of VNS treatment; nevertheless, animals underwent further rehabilitation [179]. Two studies each observed that VNS attenuated the development of brain edema [182, 183], that it is effective for the treatment of cognitive [177, 178] or motor impairments [178, 179], had neuroprotective effects [183, 184], and promoted wakefulness after TBI [180, 181].

Discussion

Transcranial magnetic stimulation

Experiments with TMS in preclinical models of TBI attracted interest rather recently with the oldest study dating back to 2015. All of the analyzed TMS studies in this scope employ rTMS protocols for effective treatment. Given that the early phases after TBI are associated with cortical hypoexcitability [185, 186], high frequency rTMS has been the major focus of interest in the studied publications. This is in line with the treatment window in these studies, which often starts relatively soon after TBI. On the other hand, low frequency rTMS induces inhibitory effects, rendering neurons less likely to fire [82], and is mostly utilized in epilepsy research [187]. It should be noted that post-TBI hyperexcitability is also observed, though after some time with an onset after approximately 2 months in preclinical models [188] and it is associated with trauma-induced epilepsy. Notwithstanding, preclinical experiments with low frequency rTMS for the prevention of TBI-induced epileptogenesis are currently quite underrated and further research is needed.

The inclusion of appropriate control groups in TMS studies deserves critical emphasis. Verdugo-Diaz et al., for instance, showed that movement restriction alone, which is necessary for stimulation in awake animals, significantly reduced post-traumatic bleeding and mortality, and improved neurobehavioral scores to the same extent observed in the rTMS group [157]. Similarly, combination of rTMS with environmental enrichment (EE) reportedly led to improvements in sensorimotor function lasting up to 6 weeks compared to the rTMS alone [154]. However, in this study rTMS was applied for only 1 week post-TBI, whilst EE lasted for 6 weeks. Unfortunately, both untreated TBI and TBI + EE controls were not included in the beamwalk tests, leaving the question unanswered whether rTMS itself had any long-term contributions to the observed improvement.

Large variabilities in the used stimulation frequencies (2–40 Hz), stimulation durations (3–20 min), treatment periods (a few days to 4 weeks), as well as heterogeneity in the used protocols for pulse trains, make a direct comparison between these studies difficult. Stimulation parameters were either taken from previous studies investigating modes of injury other than TBI [152, 155,156,157,158], from clinical studies [153], or the choice of parameters was not mentioned [138, 154]. No two studies utilize comparable stimulation intensities, thus, a correlation of the stimulation parameters to different outcomes is hindered. Nevertheless, several studies with different TBI models, namely weight drop and controlled cortical impact (CCI), showed functional improvements upon rTMS starting 1 or 2 days after TBI, [138, 154,155,156,157] with daily sessions usually administered for 1 week or longer. However, in a rat model of lateral fluid percussion injury (FPI), rTMS starting 4 days after induction of severe TBI did not show any improvements in motor behavioral outcome [152], whilst in a CCI model of pediatric TBI beneficial effects were reported after starting rTMS 9 days post-injury [153]. Similar improvements in neurological scores were also reported after moderate TBI using Feeney’s weight drop model, when rTMS was started 4 days post-injury [158]. Reported cellular and molecular biological readouts suggest that the observed functional improvements could be the result of neuroprotection, thus a critical time window for the treatment after TBI can be presumed. However, the existence of such a therapeutic window, and whether it is influenced by factors such as age, gender, and trauma severity, is unclear due to the limited number of published preclinical studies on this topic as well as the large variability in used parameters and treatment regimens.

Biological correlates of observed functional improvements could include mitigation of apoptotic signaling and cell death [152, 156], as well as reduced loss of mature neurons [155, 156] and astroglial activation [155] together with increases in cell proliferation and neurogenesis in the neurogenic niches such as the subventricular zone of lateral ventricles [156]. Moreover, upregulations in the expression levels of brain-derived neurotrophic factor (BDNF), tropomyosin receptor kinase B (TrkB, neurotrophin receptor), N-methyl-d-aspartate receptor 1 (NMDAR1, glutamate receptor) and phosphorylation of cyclic AMP response element binding protein (CREB; induced by neuronal activation) [158] support the presumption that restoration of cortical excitability early after TBI has a critical role not only in attenuation of delayed loss of cells that survived the primary impact, but also in the enhancement of regenerative responses. These results are of peculiar importance for a better understanding of underlying biological correlates of improvements that were detected in clinical applications, as most of these readouts are devoid of any possibility of direct assessment in the clinical practice. Whilst the positive results are per se encouraging—despite large variabilities in injury type, trauma severity and stimulation parameters—the translational value of preclinical studies is invariably dependent on their power in delineating correlative and causative relations between the applied stimulation parameters and observed biological readouts. Therefore, maturation of preclinical research on post-TBI rTMS from the current exploratory phase towards standardized procedures that allow for systematic comparisons is highly desirable.

Transcranial direct current stimulation

Similar to TMS, tDCS has only been under investigation in preclinical studies as a possible treatment for TBI sequelae in recent years, starting in 2016. Several of the selected studies investigated the same TBI sequelae and how tDCS could be used to treat them in a preclinical setting. Two studies from the same research group assessed the effect of tDCS on microvascular cerebral blood flow (mCBF), starting in the chronic phase 3 weeks after trauma induction using either repeated or single stimulation sessions [161, 164]. They could observe a restoration of impaired cerebrovascular reactivity to hypercapnia, improved cerebral blood flow and tissue oxygenation, which is a key factor in brain metabolism associated with brain damage in the acute phase. A decrease in blood flow regulation together with decreased tissue oxygenation is suspected to cause damage in the early phase post-injury. Moreover, a chronic reduction of local brain perfusion in patients with TBI is known to cause persisting effects on brain function [189] and is thus suspected to play a crucial role in long-term outcome. An improvement in motor function and excitability could be observed in response to a single tDCS session directly after TBI induction [160] or in the subacute phase 24 h after the injury [163]. The beneficial effect of the stimulation was apparent from the day after the stimulation in both experiments and up to 12 days later, where the experiment ended. In two other studies, the improvement in motor function in response to repeated tDCS over 4 weeks in the chronic phase was examined. The first of these studies, where stimulation was started 3 days after trauma, observed a significant difference to the sham-stimulation group from day 8 to day 26 post-injury [162]. In the second study, two groups with different time points of stimulation onsets, 1 and 3 weeks after injury, were compared [161]. The group with later onset of stimulation made a better recovery than when stimulation started 1 week after TBI, showing that tDCS led to a significant decrease in neurologic impairment and an increase in motor function, memory and learning. This finding was in part supported by another study, where tDCS was started either 1 or 2 weeks post-injury and lasted for 5 consecutive days [159]. Here, the results of the Rotarod test in the 2-week group were slightly better than in the 1-week group. However, the observed improvement in spatial memory was comparable in both groups. Long-lasting or persisting effects after the end of stimulation were assessed in four studies. The first showed a beneficial effect on motor function and spatial learning directly after tDCS sessions ended, however, 2 weeks later the animals in the other groups had recovered to a similar level [159]. The second study could show a persisting effect of the stimulation 1 week after the end of the treatment [161]. The third study investigated structural brain damage in MRI 12 days after the trauma immediately after tDCS, but did not find any significant volumetric changes such as hydrocephalus or cortical thinning in either of the groups (sham, repetitive mTBI, and repetitive mTBI with tDCS). Immunohistochemistry did not show any evidence of neuronal degeneration in sham, TBI or stimulated group. Immunohistochemical study with glial fibrillary acidic protein (GFAP) stain showed a slight hypertrophy of cell bodies and a minimal extension of cell processes in both the TBI and anodal tDCS group compared to the sham group 12 days after the trauma [160]. Another study, where stimulation was applied for 7 days starting 6 weeks after TBI, found no persisting effects after stimulation was stopped [165].

These findings lead to the conclusion that tDCS decreases the time needed for recovery. From the evidence presented above, it is unclear if tDCS is able to induce persisting changes in neuronal tissue, although an improvement of motor function and tissue oxygenation could be observed over several weeks. The effect of anesthesia on the treatment outcome is not apparent in the selected studies and the choice of anesthetizing animals during stimulation is not directly correlated to the impairment under investigation. Most studies adopted their stimulation parameters from research papers that treated impairments caused by something other than TBI [159, 163, 165] and two gave no specific reasoning for their choice of stimulation parameters [160, 161] and later reused them in publications for further investigations [163, 164].

Follow-up studies could focus on investigating changes to the established stimulation protocols and how these changes affect treatment outcome, while using electrodes with standardized surface areas or adjusting the amplitude of applied currents to reach comparable current densities. The timing of stimulation onset seems to be an important factor for a better treatment outcome, however, there are no commonalities concerning the optimal time point for the start of stimulation post-injury among these studies. Cathodal stimulation is rarely used in tDCS studies, even though it was shown to be an effective treatment to decrease impulsivity and increase attention after TBI [165], and there may be additional applications for it. Most of the studies in this scope assessed the histologic changes after TBI and tDCS treatment, which could serve as a solid basis for further research into the therapeutic mechanisms of tDCS.

Deep brain stimulation

DBS first started to find use in preclinical studies about TBI treatment in 2013. The studies selected for this review used widely different stimulation protocols and time frames for each potential treatment application, which makes a comparison between them difficult. Almost half of preclinical DBS studies applied electrical stimuli continuously for 2 h [176], 1 day [175], or several weeks [167, 170, 171]. One research group initiated DBS whenever a signal measured via external urethral sphincter electromyography exceeded a certain threshold, in an attempt to enhance voiding efficiency [172,173,174]. Another group started stimulation directly before an experimental task in order to improve cognitive outcome [166, 168], while in one study stimulation was triggered every time a rodent successfully found a hidden platform in a Morris water maze test, with the goal to reinforce learning [169]. Most of the time, animals received stimulation in multiple sessions over several days [166, 168, 169, 175] or weeks [167, 170, 171], with others only applying a single session before the animals were sacrificed for further analysis [172,173,174, 176]. For the treatment of decreased arousal and disorders of consciousness, stimulation was usually initiated shortly after injury [167, 176], while treatment of bladder dysfunction started 1 week after induction of TBI [172,173,174]. Therapy of cognitive deficits was shown to be effective in the acute [167], subacute [166, 168] and chronic phases of TBI [169, 170].

Two studies used higher frequency stimulation of 100 Hz or more in the thalamic region to increase excitability in animals suffering from decreased arousal [175] or disorders of consciousness [176]. Stimulation frequencies as low as 7.7–8 Hz were applied in the midbrain or medial septal nucleus to treat cognitive deficits [166,167,168], while 30 Hz stimulation in the lateral cerebellar nucleus was used for a similar purpose [170, 171]. All three studies investigating DBS as a treatment for bladder dysfunction in this scope originate from the same research group and used identical stimulation parameters [172,173,174]. Their triggered approach consists of 10 s of 50 Hz stimulation at amplitudes between 1 and 2.5 V. In their most recent study [174], they explored simulation of the pedunculopontine tegmental nucleus instead of the rostral pontine reticular nucleus to investigate its neural connectivity with bladder function, resulting in a similar outcome. Aronson et al. applied 130 Hz biphasic pulses in trains of 500 ms in the nucleus accumbens, whenever an animal succeeded a given task, leading to an improved spatial memory in TBI rats [169]. Only one group reported that they found no beneficial effects after stimulating the medial septal nucleus at a frequency of 100 Hz [168]. While one study did undocumented preliminary research to find optimal stimulation parameters [166], others adopted their parameters from previous studies on different topics [171, 172, 175, 176] or made the selection and optimization of the stimulation protocols part of their study [167, 168, 173, 174]. Jen et al. found an ideal stimulus length and frequency for effective stimulation for their purpose [172], only to continue with investigations regarding the optimal stimulation intensity in further studies [173, 174]. Only Aronson et al. do not describe how they chose the exact stimulation parameters they use, but mention that phasic stimulation in the nucleus accumbens might be able to promote neural plasticity [169].

Three studies found that DBS in various locations can be used to improve motor function after TBI [167, 171, 175] and three others observed an improvement in voiding efficiency [172,173,174]. Two studies found that DBS improved spatial working memory [166, 167] and attenuated hippocampal theta activity [166, 168]. In one study, researchers observed a mediation of anti-apoptotic and anti-inflammatory effects after DBS [171], while another confirmed that it may promote wakefulness [176]. Most studies did not investigate any persisting effects of DBS. However, one study observed that the beneficial effects of their task-matched stimulation approach on spatial memory persisted 10 days after stimulation cessation [169], and several clinical studies have shown before that DBS leads to long-lasting positive changes in connectivity [190,191,192]. Animals were usually awake during DBS, except in studies involving cystometric measurements where they were anesthetized [172,173,174].

Researchers should continue building upon the insights gained in these studies about DBS as an effective preclinical treatment for TBI sequelae to find out more about the underlying mechanisms pertaining to precise electrical stimulation of specific brain areas. It would be desirable to find a consensus about the most effective stimulation parameters and time frames for a variety of impairments by comparing the effects of small parameter changes, as it was already shown in some studies in this scope. Experiments often lasted for less than 1 week, and animals were often sacrificed directly after an experiment or shortly after stimulation was terminated, having left no room for investigations into possible long-term improvements. Since DBS is used as a long-term treatment in clinical studies [114], preclinical studies should also address the effects of long-term stimulation. It remains to be seen if different impairments with related underlying neurologic causes may be treatable with similar stimulation protocols by stimulating in different brain regions.

Vagus nerve stimulation

Compared to other stimulation modalities, the efficacy of VNS in the preclinical treatment of TBI sequelae has been investigated for a much longer time since 2005. Almost all VNS studies in this scope stimulated the left vagus nerve at the cervical level, except for one that targeted the right vagus nerve [183]. This consensus may stem from the fact that the right vagus nerve has more projections to the sinoatrial node of the cardiac atria and stimulation could therefore have an undesirable effect on the cardiac rhythm [130].

In the analyzed publications, most research was focused on treatment with multiple repeated stimulation sessions [177,178,179, 182, 184], while a few publications report the effects of single session VNS treatment [180, 181, 183]. The onset of the treatment in studies using repeated stimulation varied between 2 h [177, 182], 24 h [178, 184] and up to 9 days [179] after sustaining TBI. These time points correspond to different phases of post-injury pathology: early acute phase, subacute phase and chronic phase. In a clinical setup, therapy can be implemented at any point after TBI; however, early interventions are known to lead to better functional and psychological outcomes in patients [193,194,195,196]. Additionally, the long-term study of Pruitt et al. delivered stimuli within 45 ms after each successful pull trial [179], which should lead to strengthening of synaptic connections according to the STDP model of plasticity. In the studied publications, an early onset of the VNS treatment led to a faster recovery of motor skills, which is usually observed around day 2 [177, 182], as compared to a subacute onset from day 4 on [178]. Conversely, starting stimulation 24 h after TBI encouraged a faster improvement of cognitive functions; 13 days for early-onset [177] compared to 11 days for the later-onset study [178]. The study implementing VNS in the chronic phase also showed a positive effect of VNS on motor recovery [179]. However, it was sustained for 5 weeks and accompanied by physical training of the animals. Persisting effects of VNS were described for 1 week after cessation of the treatment. Multiple stimulation sessions also proved to have a neuroprotective effect on GABAergic neurons [184] and limit edema formation in the ipsilateral cortex [182]. In summary, repeated VNS aids in functional recovery after TBI and to some extent helps in constraining secondary damage.

Single stimulation after TBI led to a wake-promoting effect in free-fall injury animal models [180, 181] and the promotion of anti-inflammatory cytokine modulation with lower edema formation in a blast injury model [183]. This might indicate that an isolated VNS session could be advantageous in early post-injury stages and might lead to diminishing secondary injury. Nevertheless, clinical application of single VNS treatment would be plausible in the form of non-invasive stimulation, rather than during surgery. Transcutaneous VNS has already been proven feasible and was well tolerated in humans with severe TBI [197]. Pre-clinical studies employing this kind of VNS treatment for TBI are not available at this moment, but have been described for models of inflammation [198, 199], ischemia [200] and seizures [201].

Unlike in the case of TMS, the VNS studies in this scope use comparable stimulation protocols. Four publications coming from one research group report using the same stimulation parameters, which authors described that they were adapted from a previous study [177, 178, 182, 184]. This leads to a better reproducibility of the experiments and facilitates comparability of the results. Other studies mention implementing the same stimulation protocols as previous publications where the respective impairments had a different underlying cause than TBI [179,180,181], while one study does not mention how stimulation parameters were selected [183].

Since VNS is an established method and was FDA-approved for drug-resistant epilepsy and depression [202], there is an abundance of commercially available devices for human patients. However, similar devices for rats are currently not accessible and only some of the publications [177, 178, 182, 183] describe the electrodes they were implanting. Post-experimental re-testing of the electrodes is reported in only three of them [177, 178, 184]. None of the studies mentions pre-surgical evaluation of the devices, which might be crucial to ensure proper functionality. Similarly, observed side effects were also not reported in any of the analyzed publications, which could lead to insights into safety of VNS application in patients.

Since VNS is usually used as a long-term treatment in awake patients, the effect of anesthesia on the stimulation outcome is not investigated, unless it is specifically used as a treatment for disorders of consciousness [180, 181]. However, these studies report usage of a chloral hydrate, a drug considered not suitable for anesthesia of laboratory animals [203], and mention inducing anesthesia three times during 1 day in some of the experimental groups, which is a considerable burden for animals. Therefore, these results should be interpreted with caution.

Compared to the other stimulation methods presented above, there is more consensus between different VNS studies. This method proved to be advantageous for therapy of different conditions associated with TBI, regardless of the temporal window of its implementation and the amount of stimulation sessions. Further studies aimed at different modalities of VNS, e.g. transcutaneous VNS, and combination with other therapeutic agents, such as physiotherapy and pharmacotherapy, as well as life-long studies might lead to additional insights into better applications of VNS in humans.

Comparison between different methods

All of the methods discussed here can be used to treat motor and cognitive dysfunctions and lead to significant improvements in TBI animal models [138, 153,154,155, 159, 160, 162, 163, 166,167,168,169,170,171, 177,178,179]. Only one study found that their TMS protocol did not induce any beneficial effects regarding motor improvements [152], which was likely due to the relatively short stimulation duration they used compared to other studies. At the same time, there are a variety of other TBI sequelae that benefit from treatment with different electrical stimulation modalities. Neuroprotective effects can be induced with TMS, DBS and VNS to prevent further cell-death after injury [153, 155, 171, 183, 184]. Both TMS and tDCS are able to modulate cortical excitability leading to plasticity and increased brain activity [153, 154, 160, 165]. Suppression of cortical excitability can be achieved with TMS and tDCS as well, leading to a decrease in hyperactivity and impulsivity in animals [153, 165]. After stimulation with TMS and DBS, researchers discovered beneficial changes in histological assessment [157, 158, 168,169,170,171], while some tDCS and VNS studies show positive effects on protein expressions after treatment [159, 160, 162, 180, 183]. Finally, the studies in the scope of this review show that tDCS, DBS and VNS may effectively be used to promote wakefulness and treat disorders of consciousness caused by TBI [160, 175, 176, 180, 181].

While these stimulation methods have many treatment opportunities in common, each of them have possible applications that have not yet been observed with the other modalities in TBI animal models, giving them a status as some sort of “specialization”. TMS has been used to improve brain metabolism and potentially induce cell proliferation and neurogenesis [156], while tDCS studies showed that it can be used to increase microvascular flow and tissue oxygenation [161, 164]. This likely stems from the fact that these two methods both activate large parts of the cortex, therefore having a higher impact on the metabolism and oxygenation of the brain. Exclusively, DBS studies explored the application of electrical stimulation to improve voiding efficiency in animals with bladder dysfunction [172,173,174], since DBS can be used to specifically target diseases whose etiology is connected to single brain regions. Only VNS, which is known to decrease the disruption of the blood–brain barrier [120], has been used in preclinical studies to attenuate the development of cerebral edema after TBI [182, 183].

Translatability of the results

There are several aspects of pre-clinical studies that should be taken into consideration while analyzing their translatability into a clinical environment. Among them worth mentioning are: the relevance of the animal model, appropriate treatment, the temporal window, and side effects.

All of the analyzed studies were performed with well-established mammalian model species: rat, mouse and rabbit. The most frequently used model species was rat (28/34), with Sprague–Dawley as a leading strain (19/28), followed by Long Evans rats (7/28) and a single instance of Wistar rats used. A minor portion of analyzed studies was performed on mice (5/34) and only one publication reports experiments on New Zealand rabbits. The dominance of the rat model stems from a relatively big size of the brain in these animals, as compared to mice. This translates to convenience during surgery, especially when small electrodes are implanted, but is also important for a better spatial resolution when targeting specific brain regions [204], for instance with TMS. Common use of Sprague–Dawley rats ascertains comparability of the results within and between neurostimulation methods. However, Sprague–Dawley rats were reported to reach a faster motor skill recovery as compared to Long Evans rats [205]. Therefore, caution is recommended when comparing the two strains with each other. Moreover, all of those species are lissencephalic and display different geometry, craniospinal angle and grey-to-white matter ratio than humans [206], which is a further limitation of the translatability of results to human patients. Only one publication [153] used juvenile rats as a model for a TMS study. Since TBI is the disease with one of the highest incidences in children and youth below 19 years old [207], it is of utmost importance to further encourage studies employing neurostimulation methods as a post-traumatic therapy in young animals, with special focus on non-invasive methods.

Sex-dependent differences in the outcomes of TBI pre-clinical studies have been widely reported and reviewed in multiple studies [208,209,210]. In general, animal studies report better outcomes in females than in males, which might stem from the neuroprotective effects of estrogen and progesterone [208, 209]. The desire to determine treatment efficacy independent of hormonal status leads to the selective inclusion of males in pre-clinical studies, unless the study is specifically designed to address the sex difference itself [208, 209]. Likewise, only a small proportion of publications analyzed here reports use of female animals (2/34) [172, 179] or both sexes (2/34) [176, 180], which restrains the translatability of the results into human patients. Inclusion of female animals in experimental post-traumatic neurostimulation research is recommended for a better representation of the clinical situation.

Methods of inducing TBI varied: 15/34 studies used a weight-drop method, 10/34 fluid percussion injury, 8/34 controlled cortical impact and 1/34 performed blast injury. Except for blast injury, which is not fully consistent, these models are highly standardized and cover different types of injury, from focal to diffuse and mixed [206], corresponding to lesion diversity in patients who have survived head injury. The reported severity of the TBI model also varies: mild TBI was induced in 3/34 of studies, mild-moderate TBI in 2/34, moderate TBI in 13/34 and severe in 5/34. This does not fully mirror the clinical situation in humans, where approximately 80% of TBI is categorized as mild [211]. Nevertheless, moderate and severe TBI constitute approximately 50% of hospitalizations [212] and lead to higher mortality [213, 214]. Regrettably, a significant proportion of studies (11/34) does not report the severity level of the injury, substantially limiting their translatability. It is also worth noting that, due to anatomical and coil size differences, TMS may be able to stimulate deeper brain regions in small rodents that could otherwise not be effectively targeted in human patients [25].

Appropriate treatment requires a suitable method and stimulation protocol for the disability under investigation in the respective study. This is especially an issue for the clinical applicability of the TMS, tDCS and VNS studies in the scope of this review, since they use widely different stimulation parameters and time frames, even for the treatment of similar disabilities. In case of VNS, stimulation protocols were comparable; however, studies performing acute VNS intraoperatively might not be as clinically relevant.

The temporal window of applied stimulation methods varies highly. Early onset of the stimulation protocol was reported in 9/34 of publications analyzed, subacute in 13/34 and chronic in 13/34 of studies. Interestingly, individual methods seem to be applied at specific time points: TMS almost exclusively in the subacute stage, tDCS mostly in the subacute and chronic stages, and VNS in the acute and subacute stages, while only DBS finds application in all stages after TBI. This distribution of the time points may correspond well to the clinical situation, in which onset and duration of the therapy are highly variable [215,216,217].

Finally, possible adverse effects of the treatment are an important factor as well. The presence of side effects during pre-clinical studies might indicate plausible future problems in the clinical setting and should not be underestimated. Review articles on side effects caused by therapeutic application of TMS, tDCS and VNS in a clinical setting report only mild side effects [218,219,220], while adverse effects of DBS require more investigation in closer cooperation of scientists and clinicians [221] and are prone to bias [222]. Very few of the studies in this scope investigated possible side effects of any of these four stimulation methods and not a single one reported that they found any negative implications, which similarly hints to a possible bias and would be an important aspect in further research on this topic.

Conclusion

This literature review was conducted in order to give a comprehensive overview on the most commonly applied electrical stimulation techniques used in conjunction with preclinical models to investigate their potential for rehabilitation after TBI. Our approach focused on the specific stimulation parameters and time frames used in the analyzed studies with the goal to help optimize treatment applications. One limitation of this review is the fact that it focuses specifically on the treatment of TBI sequelae, leaving it blind to stimulation protocols used for similar impairments with different underlying causes. Nevertheless, TBI treatment is one of the main applications for electrical stimulation paradigms, which is why this review showcases a large portion of the research conducted in this field.

We found that for some stimulation methods, specifically tDCS and VNS, researchers have started using comparable protocols over the recent years, increasing their focus on the specific cellular mechanisms leading to an improved outcome. TMS and DBS, however, are used for the treatment of a diverse group of TBI sequelae, employing widely different stimulation parameters and starting at various time points after injury. This makes it difficult to find optimal treatment solutions and leaves a lot of questions about further improvements that could be achieved through small adjustments to these parameters and time frames. Further research in this field should focus on building upon the insights documented in previous publications by using comparable experimental models and varying parameters such as stimulation frequencies, amplitudes, duration, onset after injury and how often it is repeated, while looking at cognitive and behavioral improvements, as well as beneficial changes occurring at the cellular level. Researchers should look at the long-term effects of electrical stimulation methods in TBI therapy, which were rarely investigated in the publications analyzed herein. However, it is clear that all four of the stimulation modalities in the focus of this review show promising results and have the potential to shape the future of clinical treatment of patients following TBI.

Availability of data and materials

All data generated or analyzed during this study are included in this published article. Datasets with more detailed information are available from the corresponding author on reasonable request.

Abbreviations

BDNF:

Brain-derived neurotrophic factor

CaMKII:

Calcium/calmodulin-dependent protein kinase 2

cAMP:

Cyclic adenosine monophosphate

CCI:

Controlled cortical impact

CLOCK:

Circadian locomotor output cycles protein kaput

CMG:

Cystometrogram

CREB:

CAMP response element-binding protein

CRY2:

Cryptochrome 2

CST:

Corticospinal tract

CT:

Computed tomography

DBS:

Deep brain stimulation

DoC:

Disorders of consciousness

ECS:

Electrical cortical stimulation

EE:

Environmental enrichment

EEG:

Electroencephalogram

EES:

Epidural electrical stimulation

ELISA:

Enzyme-linked immunoassay

ES:

Electrical stimulation

EUS-EMG:

External urethral sphincter electromyography

fMRI:

Functional magnetic resonance imaging

FPI:

Fluid percussion injury

GABA:

Gamma-aminobutyric acid

GABABR:

Gamma-aminobutyric acid beta receptor

GAD:

Glutamic acid decarboxylase

GAP43:

Growth associated protein 43

GFAP:

Glial fibrillary acidic protein

IL-10:

Interleukin-10

IL-1β:

Interleukin-1 beta

LCN:

Lateral cerebellar nucleus

LFMS:

Low-field magnetic stimulation

LFP:

Local field potential

LHA:

Lateral hypothalamic area

LSCI:

Laser speckle contrast imaging

LSM:

Laser scanning microscopy

LTD:

Long-term depression

LTP:

Long-term potentiation

M1:

Primary motor cortex

mCBF:

Microvascular cerebral blood flow

MEP:

Motor-evoked potential

MeSH:

Medical subject headings

mNSS:

Modified Neurological Severity Score

MR:

Midbrain raphe

MRI:

Magnetic resonance imaging

mRNA:

Messenger RNA

MRS:

Magnetic resonance spectroscopy

MSN:

Medial septal nucleus

mTBI:

Mild traumatic brain injury

MUA:

Multi-unit activity

NADH:

Nicotinamide adenine dinucleotide

NeuN:

Neuronal nuclei

NMDA:

N-Methyl-d-aspartate

NSS:

Neurological severity screen

OX1R:

Orexins receptor type 1

P-CREB:

Phosphorylated CREB

p75NTR:

P75 neurotrophin receptor

PCR:

Polymerase chain reaction

PET:

Positron emission tomography

PnO:

Rostral pontine reticular nucleus

PPTg:

Pedunculopontine tegmental nucleus

PrPc:

Prion protein

PSD-95:

Postsynaptic density protein 95

RMT:

Resting motor threshold

rmTBI:

Repetitive mild traumatic brain injury

RNA:

Ribonucleic acid

RRT:

Rotarod test

rTMS:

Repetitive transcranial magnetic stimulation

RT-PCR:

Reverse transcription PCR

SEP:

Somatosensory-evoked potential

SPRT:

Single-pellet reaching task

STDP:

Spike timing-dependent plasticity

SVZ:

Subventricular zone

SYN:

Synaptophysin

TBI:

Traumatic brain injury

tDCS:

Transcranial direct current stimulation

TEM:

Transmission electron microscopy

TMS:

Transcranial magnetic stimulation

TNF-α:

Tumor necrosis factor alpha

TrkB:

Tropomyosin receptor kinase B

VNS:

Vagus nerve stimulation

α1-AR:

Alpha-1 adrenoceptor

References

  1. TBI Data | Concussion | Traumatic Brain Injury | CDC Injury Center [Internet]. 2022 [cited 2022 Oct 25]. Available from: https://www.cdc.gov/traumaticbraininjury/data/index.html.

  2. Maas AIR, Menon DK, Adelson PD, Andelic N, Bell MJ, Belli A, et al. Traumatic brain injury: integrated approaches to improve prevention, clinical care, and research. Lancet Neurol. 2017;16(12):987–1048.

    Article  PubMed  Google Scholar 

  3. Gennarelli TA, Champion HR, Sacco WJ, Copes WS, Alves WM. Mortality of patients with head injury and extracranial injury treated in trauma centers. J Trauma. 1989;29(9):1193–201 (discussion 1201–1202).

    Article  CAS  PubMed  Google Scholar 

  4. Lu J, Marmarou A, Choi S, Maas A, Murray G, Steyerberg EW, et al. Mortality from traumatic brain injury. Acta Neurochir Suppl. 2005;95:281–5.

    Article  CAS  PubMed  Google Scholar 

  5. McAllister TW. Neurobehavioral sequelae of traumatic brain injury: evaluation and management. World Psychiatry. 2008;7(1):3–10.

    Article  PubMed  PubMed Central  Google Scholar 

  6. Golding EM. Sequelae following traumatic brain injury. The cerebrovascular perspective. Brain Res Brain Res Rev. 2002;38(3):377–88.

    Article  PubMed  Google Scholar 

  7. Maloney-Wilensky E, Gracias V, Itkin A, Hoffman K, Bloom S, Yang W, et al. Brain tissue oxygen and outcome after severe traumatic brain injury: a systematic review. Crit Care Med. 2009;37(6):2057–63.

    Article  PubMed  Google Scholar 

  8. Palmer AM, Marion DW, Botscheller ML, Swedlow PE, Styren SD, DeKosky ST. Traumatic brain injury-induced excitotoxicity assessed in a controlled cortical impact model. J Neurochem. 1993;61(6):2015–24.

    Article  CAS  PubMed  Google Scholar 

  9. Yi JH, Hazell AS. Excitotoxic mechanisms and the role of astrocytic glutamate transporters in traumatic brain injury. Neurochem Int. 2006;48(5):394–403.

    Article  CAS  PubMed  Google Scholar 

  10. Shlosberg D, Benifla M, Kaufer D, Friedman A. Blood–brain barrier breakdown as a therapeutic target in traumatic brain injury. Nat Rev Neurol. 2010;6(7):393–403.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Unterberg AW, Stover J, Kress B, Kiening KL. Edema and brain trauma. Neuroscience. 2004;129(4):1021–9.

    Article  CAS  PubMed  Google Scholar 

  12. Donkin JJ, Vink R. Mechanisms of cerebral edema in traumatic brain injury: therapeutic developments. Curr Opin Neurol. 2010;23(3):293–9.

    Article  CAS  PubMed  Google Scholar 

  13. Lewén A, Matz P, Chan PH. Free radical pathways in CNS injury. J Neurotrauma. 2000;17(10):871–90.

    Article  PubMed  Google Scholar 

  14. Abdul-Muneer PM, Schuetz H, Wang F, Skotak M, Jones J, Gorantla S, et al. Induction of oxidative and nitrosative damage leads to cerebrovascular inflammation in an animal model of mild traumatic brain injury induced by primary blast. Free Radic Biol Med. 2013;60:282–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Schmidt OI, Heyde CE, Ertel W, Stahel PF. Closed head injury—an inflammatory disease? Brain Res Brain Res Rev. 2005;48(2):388–99.

    Article  PubMed  Google Scholar 

  16. Morganti-Kossmann MC, Satgunaseelan L, Bye N, Kossmann T. Modulation of immune response by head injury. Injury. 2007;38(12):1392–400.

    Article  PubMed  Google Scholar 

  17. Thomale UW, Schaser K, Kroppenstedt SN, Unterberg AW, Stover JF. Cortical hypoperfusion precedes hyperperfusion following controlled cortical impact injury. Acta Neurochir Suppl. 2002;81:229–31.

    CAS  PubMed  Google Scholar 

  18. Cheng G, Kong RH, Zhang LM, Zhang JN. Mitochondria in traumatic brain injury and mitochondrial-targeted multipotential therapeutic strategies. Br J Pharmacol. 2012;167(4):699–719.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Kurland D, Hong C, Aarabi B, Gerzanich V, Simard JM. Hemorrhagic progression of a contusion after traumatic brain injury: a review. J Neurotrauma. 2012;29(1):19–31.

    Article  PubMed  PubMed Central  Google Scholar 

  20. Kelly DF, Kordestani RK, Martin NA, Nguyen T, Hovda DA, Bergsneider M, et al. Hyperemia following traumatic brain injury: relationship to intracranial hypertension and outcome. J Neurosurg. 1996;85(5):762–71.

    Article  CAS  PubMed  Google Scholar 

  21. Witcher KG, Bray CE, Chunchai T, Zhao F, O’Neil SM, Gordillo AJ, et al. Traumatic brain injury causes chronic cortical inflammation and neuronal dysfunction mediated by microglia. J Neurosci. 2021;41(7):1597–616.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Wang KK, Yang Z, Zhu T, Shi Y, Rubenstein R, Tyndall JA, et al. An update on diagnostic and prognostic biomarkers for traumatic brain injury. Expert Rev Mol Diagn. 2018;18(2):165–80.

    Article  PubMed  PubMed Central  Google Scholar 

  23. Algattas H, Huang JH. Traumatic brain injury pathophysiology and treatments: early, intermediate, and late phases post-injury. Int J Mol Sci. 2013;15(1):309–41.

    Article  PubMed  PubMed Central  Google Scholar 

  24. Villamar MF, Santos Portilla A, Fregni F, Zafonte R. Noninvasive brain stimulation to modulate neuroplasticity in traumatic brain injury. Neuromodulation. 2012;15(4):326–38.

    Article  PubMed  Google Scholar 

  25. Surendrakumar S, Rabelo TK, Campos ACP, Mollica A, Abrahao A, Lipsman N, et al. Neuromodulation therapies in pre-clinical models of traumatic brain injury: systematic review and translational applications. J Neurotrauma. 2022. https://doi.org/10.1089/neu.2022.0286.

    Article  PubMed  Google Scholar 

  26. Iglesias AH. Transcranial magnetic stimulation as treatment in multiple neurologic conditions. Curr Neurol Neurosci Rep. 2020;20(1):1.

    Article  PubMed  Google Scholar 

  27. Ni Z, Kim SJ, Phielipp N, Ghosh S, Udupa K, Gunraj CA, et al. Pallidal deep brain stimulation modulates cortical excitability and plasticity. Ann Neurol. 2018;83(2):352–62.

    Article  PubMed  Google Scholar 

  28. Narapareddy BR, Narapareddy L, Lin A, Wigh S, Nanavati J, Dougherty J, et al. Treatment of depression after traumatic brain injury: a systematic review focused on pharmacological and neuromodulatory interventions. Psychosomatics. 2020;61(5):481–97.

    Article  PubMed  Google Scholar 

  29. Tsai PY, Chen YC, Wang JY, Chung KH, Lai CH. Effect of repetitive transcranial magnetic stimulation on depression and cognition in individuals with traumatic brain injury: a systematic review and meta-analysis. Sci Rep. 2021;11(1):16940.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Nadim F, Bucher D. Neuromodulation of neurons and synapses. Curr Opin Neurobiol. 2014;29:48–56.

    Article  CAS  PubMed  Google Scholar 

  31. Skarpaas TL, Jarosiewicz B, Morrell MJ. Brain-responsive neurostimulation for epilepsy (RNS® System). Epilepsy Res. 2019;153:68–70.

    Article  PubMed  Google Scholar 

  32. Schuepbach WMM, Rau J, Knudsen K, Volkmann J, Krack P, Timmermann L, et al. Neurostimulation for Parkinson’s disease with early motor complications. N Engl J Med. 2013;368(7):610–22.

    Article  CAS  PubMed  Google Scholar 

  33. Urgun K, Chan A, Sahyouni R, Tran KD, Hsu F, Vadera S. Application of responsive neurostimulation as both a diagnostic tool for seizure localization and a supplementary tool for surgical management in patients with multiple epileptogenic foci: a case series. Turk Neurosurg. 2021. https://doi.org/10.5137/1019-5149.JTN.32629-20.2.

    Article  Google Scholar 

  34. Plow EB, Machado A. Invasive neurostimulation in stroke rehabilitation. Neurotherapeutics. 2014;11(3):572–82.

    Article  PubMed  Google Scholar 

  35. Merrill DR, Bikson M, Jefferys JGR. Electrical stimulation of excitable tissue: design of efficacious and safe protocols. J Neurosci Methods. 2005;141(2):171–98.

    Article  PubMed  Google Scholar 

  36. Idlett S, Halder M, Zhang T, Quevedo J, Brill N, Gu W, et al. Assessment of axonal recruitment using model-guided preclinical spinal cord stimulation in the ex vivo adult mouse spinal cord. J Neurophysiol. 2019;122(4):1406–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Knorr S, Musacchio T, Paulat R, Matthies C, Endres H, Wenger N, et al. Experimental deep brain stimulation in rodent models of movement disorders. Exp Neurol. 2022;348: 113926.

    Article  PubMed  Google Scholar 

  38. Ranck JB. Which elements are excited in electrical stimulation of mammalian central nervous system: a review. Brain Res. 1975;98(3):417–40.

    Article  PubMed  Google Scholar 

  39. Kirsch AD, Hassin-Baer S, Matthies C, Volkmann J, Steigerwald F. Anodic versus cathodic neurostimulation of the subthalamic nucleus: a randomized-controlled study of acute clinical effects. Parkinsonism Relat Disord. 2018;55:61–7.

    Article  PubMed  Google Scholar 

  40. BeMent SL, Ranck JB. A quantitative study of electrical stimulation of central myelinated fibers. Exp Neurol. 1969;24(2):147–70.

    Article  CAS  PubMed  Google Scholar 

  41. Burke D, Kiernan MC, Bostock H. Excitability of human axons. Clin Neurophysiol. 2001;112(9):1575–85.

    Article  CAS  PubMed  Google Scholar 

  42. Safronov BV, Wolff M, Vogel W. Excitability of the soma in central nervous system neurons. Biophys J. 2000;78(6):2998–3010.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Rattay F. The basic mechanism for the electrical stimulation of the nervous system. Neuroscience. 1999;89(2):335–46.

    Article  CAS  PubMed  Google Scholar 

  44. Waxman SG, Ritchie JM. Molecular dissection of the myelinated axon. Ann Neurol. 1993;33(2):121–36.

    Article  CAS  PubMed  Google Scholar 

  45. Trigo FF. Antidromic analog signaling. Front Cell Neurosci. 2019. https://doi.org/10.3389/fncel.2019.00354.

    Article  PubMed  PubMed Central  Google Scholar 

  46. Mateus JC, Lopes C, Aroso M, Costa AR, Gerós A, Meneses J, et al. Bidirectional flow of action potentials in axons drives activity dynamics in neuronal cultures. J Neural Eng. 2021;18(6):066045.

    Article  Google Scholar 

  47. Sasaki T. The axon as a unique computational unit in neurons. Neurosci Res. 2013;75(2):83–8.

    Article  CAS  PubMed  Google Scholar 

  48. Rama S, Zbili M, Debanne D. Signal propagation along the axon. Curr Opin Neurobiol. 2018;51:37–44.

    Article  CAS  PubMed  Google Scholar 

  49. Brocker DT, Grill WM. Principles of electrical stimulation of neural tissue. Handb Clin Neurol. 2013;116:3–18.

    Article  PubMed  Google Scholar 

  50. Lafon B, Rahman A, Bikson M, Parra LC. Direct current stimulation alters neuronal input/output function. Brain Stimul. 2017;10(1):36–45.

    Article  PubMed  Google Scholar 

  51. Gärtner A, Staiger V. Neurotrophin secretion from hippocampal neurons evoked by long-term-potentiation-inducing electrical stimulation patterns. Proc Natl Acad Sci USA. 2002;99(9):6386–91.

    Article  PubMed  PubMed Central  Google Scholar 

  52. Esmaeilpour K, Sheibani V, Shabani M, Mirnajafi-Zadeh J. Low frequency electrical stimulation has time dependent improving effect on kindling-induced impairment in long-term potentiation in rats. Brain Res. 2017;1668:20–7.

    Article  CAS  PubMed  Google Scholar 

  53. Thoenen H. Neurotrophins and activity-dependent plasticity. In: Progress in brain research (neural plasticity and regeneration), vol 128. Elsevier; 2000, p. 183–91. Available from: https://www.sciencedirect.com/science/article/pii/S0079612300280163.

  54. Fauth M, Tetzlaff C. Opposing effects of neuronal activity on structural plasticity. Front Neuroanat. 2016. https://doi.org/10.3389/fnana.2016.00075.

    Article  PubMed  PubMed Central  Google Scholar 

  55. Lynch MA. Long-term potentiation and memory. Physiol Rev. 2004;84(1):87–136.

    Article  CAS  PubMed  Google Scholar 

  56. Albensi BC, Janigro D. Traumatic brain injury and its effects on synaptic plasticity. Brain Inj. 2003;17(8):653–63.

    Article  PubMed  Google Scholar 

  57. Albensi BC, Oliver DR, Toupin J, Odero G. Electrical stimulation protocols for hippocampal synaptic plasticity and neuronal hyper-excitability: are they effective or relevant? Exp Neurol. 2007;204(1):1–13.

    Article  PubMed  Google Scholar 

  58. Brown GP, Blitzer RD, Connor JH, Wong T, Shenolikar S, Iyengar R, et al. Long-term potentiation induced by theta frequency stimulation is regulated by a protein phosphatase-1-operated gate. J Neurosci. 2000;20(21):7880–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Zakharenko SS, Zablow L, Siegelbaum SA. Visualization of changes in presynaptic function during long-term synaptic plasticity. Nat Neurosci. 2001;4(7):711–7.

    Article  CAS  PubMed  Google Scholar 

  60. Bear MF, Malenka RC. Synaptic plasticity: LTP and LTD. Curr Opin Neurobiol. 1994;4(3):389–99.

    Article  CAS  PubMed  Google Scholar 

  61. Ito M. Long-term depression. Annu Rev Neurosci. 1989;12:85–102.

    Article  CAS  PubMed  Google Scholar 

  62. O’Neil DA, Nicholas MA, Lajud N, Kline AE, Bondi CO. Preclinical models of traumatic brain injury: emerging role of glutamate in the pathophysiology of depression. Front Pharmacol. 2018;9:579.

    Article  PubMed  PubMed Central  Google Scholar 

  63. Reeves TM, Lyeth BG, Povlishock JT. Long-term potentiation deficits and excitability changes following traumatic brain injury. Exp Brain Res. 1995;106(2):248–56.

    Article  CAS  PubMed  Google Scholar 

  64. Schwarzbach E, Bonislawski DP, Xiong G, Cohen AS. Mechanisms underlying the inability to induce area CA1 LTP in the mouse after traumatic brain injury. Hippocampus. 2006;16(6):541–50.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Albensi BC, Sullivan PG, Thompson MB, Scheff SW, Mattson MP. Cyclosporin ameliorates traumatic brain-injury-induced alterations of hippocampal synaptic plasticity. Exp Neurol. 2000;162(2):385–9.

    Article  CAS  PubMed  Google Scholar 

  66. Bi GQ, Rubin J. Timing in synaptic plasticity: from detection to integration. Trends Neurosci. 2005;28(5):222–8.

    Article  CAS  PubMed  Google Scholar 

  67. Sjöström PJ, Nelson SB. Spike timing, calcium signals and synaptic plasticity. Curr Opin Neurobiol. 2002;12(3):305–14.

    Article  PubMed  Google Scholar 

  68. Brzosko Z, Mierau SB, Paulsen O. Neuromodulation of spike-timing-dependent plasticity: past, present, and future. Neuron. 2019;103(4):563–81.

    Article  CAS  PubMed  Google Scholar 

  69. Dan Y, Poo MM. Spike timing-dependent plasticity of neural circuits. Neuron. 2004;44(1):23–30.

    Article  CAS  PubMed  Google Scholar 

  70. Fino E, Glowinski J, Venance L. Bidirectional activity-dependent plasticity at corticostriatal synapses. J Neurosci. 2005;25(49):11279–87.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Morera-Herreras T, Gioanni Y, Perez S, Vignoud G, Venance L. Environmental enrichment shapes striatal spike-timing-dependent plasticity in vivo. Sci Rep. 2019;9(1):19451.

    Article  PubMed  PubMed Central  Google Scholar 

  72. Nishimura Y, Perlmutter SI, Eaton RW, Fetz EE. Spike-timing-dependent plasticity in primate corticospinal connections induced during free behavior. Neuron. 2013;80(5):1301–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Jacob V, Brasier DJ, Erchova I, Feldman D, Shulz DE. Spike timing-dependent synaptic depression in the in vivo barrel cortex of the rat. J Neurosci. 2007;27(6):1271–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Itami C, Huang JY, Yamasaki M, Watanabe M, Lu HC, Kimura F. Developmental switch in spike timing-dependent plasticity and cannabinoid-dependent reorganization of the thalamocortical projection in the barrel cortex. J Neurosci. 2016;36(26):7039–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Guo Y, Zhang W, Chen X, Fu J, Cheng W, Song D, et al. Timing-dependent LTP and LTD in mouse primary visual cortex following different visual deprivation models. PLoS ONE. 2017;12(5): e0176603.

    Article  PubMed  PubMed Central  Google Scholar 

  76. Ranieri F, Coppola G, Musumeci G, Capone F, Di Pino G, Parisi V, et al. Evidence for associative plasticity in the human visual cortex. Brain Stimul. 2019;12(3):705–13.

    Article  PubMed  Google Scholar 

  77. Foysal KMR, de Carvalho F, Baker SN. Spike timing-dependent plasticity in the long-latency stretch reflex following paired stimulation from a wearable electronic device. J Neurosci. 2016;36(42):10823–30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Urbin MA, Ozdemir RA, Tazoe T, Perez MA. Spike-timing-dependent plasticity in lower-limb motoneurons after human spinal cord injury. J Neurophysiol. 2017;118(4):2171–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Camacho-Conde JA, del Rosario G-B, Carretero-Rey M, Khan ZU. Therapeutic potential of brain stimulation techniques in the treatment of mental, psychiatric, and cognitive disorders. CNS Neurosci Ther. 2023;29(1):8–23.

    Article  PubMed  Google Scholar 

  80. Polanía R, Nitsche MA, Ruff CC. Studying and modifying brain function with non-invasive brain stimulation. Nat Neurosci. 2018;21(2):174–87.

    Article  PubMed  Google Scholar 

  81. Zaninotto AL, El-Hagrassy MM, Green JR, Babo M, Paglioni VM, Benute GG, et al. Transcranial direct current stimulation (tDCS) effects on traumatic brain injury (TBI) recovery: a systematic review. Dement Neuropsychol. 2019;13:172–9.

    Article  PubMed  PubMed Central  Google Scholar 

  82. Lefaucheur JP. Chapter 37—transcranial magnetic stimulation. In: Levin KH, Chauvel P, editors. Handbook of clinical neurology (clinical neurophysiology: basis and technical aspects), vol. 160. Elsevier: Amsterdam; 2019. p. 559–80.

    Chapter  Google Scholar 

  83. Chail A, Saini RK, Bhat PS, Srivastava K, Chauhan V. Transcranial magnetic stimulation: a review of its evolution and current applications. Ind Psychiatry J. 2018;27(2):172–80.

    Article  PubMed  PubMed Central  Google Scholar 

  84. Groppa S, Oliviero A, Eisen A, Quartarone A, Cohen LG, Mall V, et al. A practical guide to diagnostic transcranial magnetic stimulation: report of an IFCN committee. Clin Neurophysiol. 2012;123(5):858–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Deng ZD, Lisanby SH, Peterchev AV. Electric field depth–focality tradeoff in transcranial magnetic stimulation: simulation comparison of 50 coil designs. Brain Stimul. 2013;6(1):1–13.

    Article  PubMed  Google Scholar 

  86. Klomjai W, Katz R, Lackmy-Vallée A. Basic principles of transcranial magnetic stimulation (TMS) and repetitive TMS (rTMS). Ann Phys Rehabil Med. 2015;58(4):208–13.

    Article  PubMed  Google Scholar 

  87. Voigt J, Carpenter L, Leuchter A. A systematic literature review of the clinical efficacy of repetitive transcranial magnetic stimulation (rTMS) in non-treatment resistant patients with major depressive disorder. BMC Psychiatry. 2019;19(1):13.

    Article  PubMed  PubMed Central  Google Scholar 

  88. Garnaat SL, Yuan S, Wang H, Philip NS, Carpenter LL. Updates on transcranial magnetic stimulation therapy for major depressive disorder. Psychiatr Clin N Am. 2018;41(3):419–31.

    Article  Google Scholar 

  89. Rehn S, Eslick GD, Brakoulias V. A meta-analysis of the effectiveness of different cortical targets used in repetitive transcranial magnetic stimulation (rTMS) for the treatment of obsessive–compulsive disorder (OCD). Psychiatr Q. 2018;89(3):645–65.

    Article  PubMed  Google Scholar 

  90. Cocchi L, Zalesky A, Nott Z, Whybird G, Fitzgerald PB, Breakspear M. Transcranial magnetic stimulation in obsessive-compulsive disorder: a focus on network mechanisms and state dependence. NeuroImage Clin. 2018;19:661–74.

    Article  PubMed  PubMed Central  Google Scholar 

  91. Khedr EM, Kotb HI, Mostafa MG, Mohamad MF, Amr SA, Ahmed MA, et al. Repetitive transcranial magnetic stimulation in neuropathic pain secondary to malignancy: a randomized clinical trial. Eur J Pain. 2015;19(4):519–27.

    Article  CAS  PubMed  Google Scholar 

  92. Attal N, Ayache SS, De Andrade DC, Mhalla A, Baudic S, Jazat F, et al. Repetitive transcranial magnetic stimulation and transcranial direct-current stimulation in neuropathic pain due to radiculopathy: a randomized sham-controlled comparative study. PAIN. 2016;157(6):1224.

    Article  PubMed  Google Scholar 

  93. Zeiler FA, Matuszczak M, Teitelbaum J, Gillman LM, Kazina CJ. Transcranial magnetic stimulation for status epilepticus. Epilepsy Res Treat. 2015;2015: e678074.

    Google Scholar 

  94. Fisicaro F, Lanza G, Grasso AA, Pennisi G, Bella R, Paulus W, et al. Repetitive transcranial magnetic stimulation in stroke rehabilitation: review of the current evidence and pitfalls. Ther Adv Neurol Disord. 2019;12:1756286419878317.

    Article  PubMed  PubMed Central  Google Scholar 

  95. Chen X, Yin L, An Y, Yan H, Zhang T, Lu X, et al. Effects of repetitive transcranial magnetic stimulation in multiple sclerosis: a systematic review and meta-analysis. Mult Scler Relat Disord. 2022;59: 103564.

    Article  PubMed  Google Scholar 

  96. Edinoff AN, Hegefeld TL, Petersen M, Patterson JC, Yossi C, Slizewski J, et al. Transcranial magnetic stimulation for post-traumatic stress disorder. Front Psychiatry. 2022;13: 701348.

    Article  PubMed  PubMed Central  Google Scholar 

  97. Chou YH, Hickey PT, Sundman M, Song AW, Chen NK. Effects of repetitive transcranial magnetic stimulation on motor symptoms in Parkinson disease: a systematic review and meta-analysis. JAMA Neurol. 2015;72(4):432–40.

    Article  PubMed  PubMed Central  Google Scholar 

  98. Alemam AI, Eltantawi MA. Repetitive transcranial magnetic stimulation in treatment of levodopa-induced dyskinesia in Parkinson’s disease. J Neurol Res. 2019;9(3):28–34.

    Article  Google Scholar 

  99. Brunoni AR, Nitsche MA, Bolognini N, Bikson M, Wagner T, Merabet L, et al. Clinical research with transcranial direct current stimulation (tDCS): challenges and future directions. Brain Stimul. 2012;5(3):175–95.

    Article  PubMed  Google Scholar 

  100. Stagg CJ, Antal A, Nitsche MA. Physiology of transcranial direct current stimulation. J ECT. 2018;34(3):144.

    Article  PubMed  Google Scholar 

  101. Ardolino G, Bossi B, Barbieri S, Priori A. Non-synaptic mechanisms underlie the after-effects of cathodal transcutaneous direct current stimulation of the human brain. J Physiol. 2005;568(Pt 2):653–63.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Chase HW, Boudewyn MA, Carter CS, Phillips ML. Transcranial direct current stimulation: a roadmap for research, from mechanism of action to clinical implementation. Mol Psychiatry. 2020;25(2):397–407.

    Article  PubMed  Google Scholar 

  103. Liu HH, He XK, Chen HY, Peng CW, Rotenberg A, Juan CH, et al. Neuromodulatory effects of transcranial direct current stimulation on motor excitability in rats. Neural Plast. 2019;2019: e4252943.

    Article  Google Scholar 

  104. Wiethoff S, Hamada M, Rothwell JC. Variability in response to transcranial direct current stimulation of the motor cortex. Brain Stimul. 2014;7(3):468–75.

    Article  PubMed  Google Scholar 

  105. Begemann MJ, Brand BA, Ćurčić-Blake B, Aleman A, Sommer IE. Efficacy of non-invasive brain stimulation on cognitive functioning in brain disorders: a meta-analysis. Psychol Med. 2020;50(15):2465–86.

    Article  PubMed  PubMed Central  Google Scholar 

  106. Fregni F, El-Hagrassy MM, Pacheco-Barrios K, Carvalho S, Leite J, Simis M, et al. Evidence-based guidelines and secondary meta-analysis for the use of transcranial direct current stimulation in neurological and psychiatric disorders. Int J Neuropsychopharmacol. 2021;24(4):256–313.

    Article  PubMed  Google Scholar 

  107. Foote KD, Seignourel P, Fernandez HH, Romrell J, Whidden E, Jacobson C, et al. Dual electrode thalamic deep brain stimulation for the treatment of posttraumatic and multiple sclerosis tremor. Neurosurgery. 2006;58(4 Suppl 2):ONS-280-285 (discussion ONS-285–286).

    Google Scholar 

  108. Krauss JK, Lipsman N, Aziz T, Boutet A, Brown P, Chang JW, et al. Technology of deep brain stimulation: current status and future directions. Nat Rev Neurol. 2021;17(2):75–87.

    Article  PubMed  Google Scholar 

  109. Kokkonen A, Honkanen EA, Corp DT, Joutsa J. Neurobiological effects of deep brain stimulation: a systematic review of molecular brain imaging studies. NeuroImage. 2022;260: 119473.

    Article  PubMed  Google Scholar 

  110. Herrington TM, Cheng JJ, Eskandar EN. Mechanisms of deep brain stimulation. J Neurophysiol. 2016;115(1):19–38.

    Article  CAS  PubMed  Google Scholar 

  111. Oluigbo CO, Salma A, Rezai AR. Deep brain stimulation for neurological disorders. IEEE Rev Biomed Eng. 2012;5:88–99.

    Article  PubMed  Google Scholar 

  112. Khan IS, D’Agostino EN, Calnan DR, Lee JE, Aronson JP. Deep brain stimulation for memory modulation: a new frontier. World Neurosurg. 2019;126:638–46.

    Article  PubMed  Google Scholar 

  113. Sullivan CRP, Olsen S, Widge AS. Deep brain stimulation for psychiatric disorders: from focal brain targets to cognitive networks. NeuroImage. 2021;225: 117515.

    Article  PubMed  Google Scholar 

  114. Kundu B, Brock AA, Englot DJ, Butson CR, Rolston JD. Deep brain stimulation for the treatment of disorders of consciousness and cognition in traumatic brain injury patients: a review. Neurosurg Focus. 2018;45(2):E14.

    Article  PubMed  PubMed Central  Google Scholar 

  115. Eljamel S. Mechanism of action and overview of vagus nerve stimulation technology. In: Eljamel S, Slavin KV, editors. Neurostimulation. Oxford: Wiley; 2013. p. 109–20. https://doi.org/10.1002/9781118346396.ch13.

    Chapter  Google Scholar 

  116. Attenello F, Amar AP, Liu C, Apuzzo MLJ. Theoretical basis of vagus nerve stimulation. Prog Neurol Surg. 2015;29:20–8.

    Article  PubMed  Google Scholar 

  117. Martlé V, Peremans K, Raedt R, Vermeire S, Vonck K, Boon P, et al. Regional brain perfusion changes during standard and microburst vagus nerve stimulation in dogs. Epilepsy Res. 2014;108(4):616–22.

    Article  PubMed  Google Scholar 

  118. Yang J, Phi JH. The present and future of vagus nerve stimulation. J Korean Neurosurg Soc. 2019;62(3):344–52.

    Article  PubMed  PubMed Central  Google Scholar 

  119. Rosso P, Iannitelli A, Pacitti F, Quartini A, Fico E, Fiore M, et al. Vagus nerve stimulation and neurotrophins: a biological psychiatric perspective. Neurosci Biobehav Rev. 2020;113:338–53.

    Article  PubMed  Google Scholar 

  120. Lopez NE, Krzyzaniak MJ, Costantini TW, Putnam J, Hageny AM, Eliceiri B, et al. Vagal nerve stimulation decreases blood–brain barrier disruption after traumatic brain injury. J Trauma Acute Care Surg. 2012;72(6):1562–6.

    Article  PubMed  Google Scholar 

  121. Kaya M, Orhan N, Karabacak E, Bahceci MB, Arican N, Ahishali B, et al. Vagus nerve stimulation inhibits seizure activity and protects blood–brain barrier integrity in kindled rats with cortical dysplasia. Life Sci. 2013;92(4):289–97.

    Article  CAS  PubMed  Google Scholar 

  122. Yang Y, Yang LY, Orban L, Cuylear D, Thompson J, Simon B, et al. Non-invasive vagus nerve stimulation reduces blood–brain barrier disruption in a rat model of ischemic stroke. Brain Stimul. 2018;11(4):689–98.

    Article  PubMed  PubMed Central  Google Scholar 

  123. Ulloa L, Quiroz-Gonzalez S, Torres-Rosas R. Nerve stimulation: immunomodulation and control of inflammation. Trends Mol Med. 2017;23(12):1103–20.

    Article  PubMed  PubMed Central  Google Scholar 

  124. Mertens A, Raedt R, Gadeyne S, Carrette E, Boon P, Vonck K. Recent advances in devices for vagus nerve stimulation. Expert Rev Med Devices. 2018;15(8):527–39.

    Article  CAS  PubMed  Google Scholar 

  125. Afra P, Adamolekun B, Aydemir S, Watson GDR. Evolution of the vagus nerve stimulation (VNS) therapy system technology for drug-resistant epilepsy. Front Med Technol. 2021. https://doi.org/10.3389/fmedt.2021.696543.

    Article  PubMed  PubMed Central  Google Scholar 

  126. O’Reardon JP, Cristancho P, Peshek AD. Vagus nerve stimulation (VNS) and treatment of depression: to the brainstem and beyond. Psychiatry Edgmont. 2006;3(5):54–63.

    PubMed  PubMed Central  Google Scholar 

  127. Jorge RE, Robinson RG, Moser D, Tateno A, Crespo-Facorro B, Arndt S. Major depression following traumatic brain injury. Arch Gen Psychiatry. 2004;61(1):42–50.

    Article  PubMed  Google Scholar 

  128. Lavoie S, Sechrist S, Quach N, Ehsanian R, Duong T, Gotlib IH, et al. Depression in men and women one year following traumatic brain injury (TBI): a TBI model systems study. Front Psychol. 2017. https://doi.org/10.3389/fpsyg.2017.00634.

    Article  PubMed  PubMed Central  Google Scholar 

  129. Larkin M, Meyer RM, Szuflita NS, Severson MA, Levine ZT. Post-traumatic, drug-resistant epilepsy and review of seizure control outcomes from blinded, randomized controlled trials of brain stimulation treatments for drug-resistant epilepsy. Cureus. 2016;8(8):e744.

  130. Noller CM, Levine YA, Urakov TM, Aronson JP, Nash MS. Vagus nerve stimulation in rodent models: an overview of technical considerations. Front Neurosci. 2019. https://doi.org/10.3389/fnins.2019.00911.

    Article  PubMed  PubMed Central  Google Scholar 

  131. Rossetti N, Hagler J, Kateb P, Cicoira F. Neural and electromyography PEDOT electrodes for invasive stimulation and recording. J Mater Chem C. 2021;9(23):7243–63.

    Article  CAS  Google Scholar 

  132. Ehlich J, Migliaccio L, Sahalianov I, Nikić M, Brodský J, Gablech I, et al. Direct measurement of oxygen reduction reactions at neurostimulation electrodes. J Neural Eng. 2022. https://doi.org/10.1088/1741-2552/ac77c0.

    Article  PubMed  Google Scholar 

  133. Qin C, Yue Z, Wallace GG, Chen J. Bipolar electrochemical stimulation using conducting polymers for wireless electroceuticals and future directions. ACS Appl Bio Mater. 2022;5(11):5041–56.

    Article  CAS  PubMed  Google Scholar 

  134. Cogan SF. Neural stimulation and recording electrodes. Annu Rev Biomed Eng. 2008;10(1):275–309.

    Article  CAS  PubMed  Google Scholar 

  135. Latchoumane CFV, Barany DA, Karumbaiah L, Singh T. Neurostimulation and reach-to-grasp function recovery following acquired brain injury: insight from pre-clinical rodent models and human applications. Front Neurol. 2020. https://doi.org/10.3389/fneur.2020.00835.

    Article  PubMed  PubMed Central  Google Scholar 

  136. Li S, Zaninotto AL, Neville IS, Paiva WS, Nunn D, Fregni F. Clinical utility of brain stimulation modalities following traumatic brain injury: current evidence. Neuropsychiatr Dis Treat. 2015;11:1573–86.

    PubMed  PubMed Central  Google Scholar 

  137. Kuo CW, Chang MY, Liu HH, He XK, Chan SY, Huang YZ, et al. Cortical electrical stimulation ameliorates traumatic brain injury-induced sensorimotor and cognitive deficits in rats. Front Neural Circuits. 2021. https://doi.org/10.3389/fncir.2021.693073.

    Article  PubMed  PubMed Central  Google Scholar 

  138. Yoon YS, Cho KH, Kim ES, Lee MS, Lee KJ. Effect of epidural electrical stimulation and repetitive transcranial magnetic stimulation in rats with diffuse traumatic brain injury. Ann Rehabil Med. 2015;39(3):416–24.

    Article  PubMed  PubMed Central  Google Scholar 

  139. Kim H, Kim HI, Kim YH, Kim SY, Shin YI. An animal study to examine the effects of the bilateral, epidural cortical stimulation on the progression of amyotrophic lateral sclerosis. J Neuroeng Rehabil. 2014;11:139.

    Article  PubMed  PubMed Central  Google Scholar 

  140. Adkins-Muir DL, Jones TA. Cortical electrical stimulation combined with rehabilitative training: enhanced functional recovery and dendritic plasticity following focal cortical ischemia in rats. Neurol Res. 2013. https://doi.org/10.1179/016164103771953853.

    Article  Google Scholar 

  141. Adachi R, Yang C. Electroconvulsive therapy for traumatic brain injury and schizoaffective disorder. Cureus. 2021;13(7):e16390.

  142. Ma R, Xia X, Zhang W, Lu Z, Wu Q, Cui J, et al. High gamma and beta temporal interference stimulation in the human motor cortex improves motor functions. Front Neurosci. 2022;15: 800436.

    Article  PubMed  PubMed Central  Google Scholar 

  143. Lee S, Park J, Choi DS, Lee C, Im CH. Multipair transcranial temporal interference stimulation for improved focalized stimulation of deep brain regions: a simulation study. Comput Biol Med. 2022;143: 105337.

    Article  PubMed  Google Scholar 

  144. Deans JK, Powell AD, Jefferys JGR. Sensitivity of coherent oscillations in rat hippocampus to AC electric fields. J Physiol. 2007;583(2):555–65.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Hutcheon B, Yarom Y. Resonance, oscillation and the intrinsic frequency preferences of neurons. Trends Neurosci. 2000;23(5):216–22.

    Article  CAS  PubMed  Google Scholar 

  146. Grossman N, Bono D, Dedic N, Kodandaramaiah SB, Rudenko A, Suk HJ, et al. Noninvasive deep brain stimulation via temporally interfering electric fields. Cell. 2017;169(6):1029-1041.e16.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Schmidt T, Jakešová M, Đerek V, Kornmueller K, Tiapko O, Bischof H, et al. Light stimulation of neurons on organic photocapacitors induces action potentials with millisecond precision. Adv Mater Technol. 2022. https://doi.org/10.1002/admt.202101159.

    Article  PubMed  PubMed Central  Google Scholar 

  148. Rand D, Jakešová M, Lubin G, Vėbraitė I, David-Pur M, Đerek V, et al. Direct electrical neurostimulation with organic pigment photocapacitors. Adv Mater Deerfield Beach Fla. 2018;30(25): e1707292.

    Article  Google Scholar 

  149. Missey F, Botzanowski B, Migliaccio L, Acerbo E, Głowacki ED, Williamson A. Organic electrolytic photocapacitors for stimulation of the mouse somatosensory cortex. J Neural Eng. 2021;18(6):066016.

    Article  Google Scholar 

  150. Missey F, Donahue MJ, Weber P, Ngom I, Acerbo E, Botzanowski B, et al. Laser-driven wireless deep brain stimulation using temporal interference and organic electrolytic photocapacitors. Adv Funct Mater. 2022;32(33):2200691.

    Article  CAS  Google Scholar 

  151. Page MJ, McKenzie J, Bossuyt P, Boutron I, Hoffmann T, Mulrow C, et al. The PRISMA 2020 statement: an updated guideline for reporting systematic reviews. MetaArXiv. 2020 [cited 2022 Jan 31]. Available from: https://osf.io/preprints/metaarxiv/v7gm2/.

  152. Yoon KJ, Lee YT, Chung PW, Lee YK, Kim DY, Chun MH. Effects of repetitive transcranial magnetic stimulation on behavioral recovery during early stage of traumatic brain injury in rats. J Korean Med Sci. 2015;30(10):1496–502.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Lu H, Kobilo T, Robertson C, Tong S, Celnik P, Pelled G. Transcranial magnetic stimulation facilitates neurorehabilitation after pediatric traumatic brain injury. Sci Rep. 2015. https://doi.org/10.1038/srep14769.

    Article  PubMed  PubMed Central  Google Scholar 

  154. Shin SS, Krishnan V, Stokes W, Robertson C, Celnik P, Chen Y, et al. Transcranial magnetic stimulation and environmental enrichment enhances cortical excitability and functional outcomes after traumatic brain injury. Brain Stimul. 2018;11(6):1306–13.

    Article  PubMed  PubMed Central  Google Scholar 

  155. Sekar S, Zhang Y, Miranzadeh Mahabadi H, Parvizi A, Taghibiglou C. Low-field magnetic stimulation restores cognitive and motor functions in the mouse model of repeated traumatic brain injury: role of cellular prion protein. J Neurotrauma. 2019;36(22):3103–14.

    Article  PubMed  Google Scholar 

  156. Lu X, Bao X, Li J, Zhang G, Guan J, Gao Y, et al. High-frequency repetitive transcranial magnetic stimulation for treating moderate traumatic brain injury in rats: a pilot study. Exp Ther Med. 2017;13(5):2247–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  157. Verdugo-Diaz L, Estrada-Rojo F, Garcia-Espinoza A, Hernandez-Lopez E, Hernandez-Chavez A, Guzman-Uribe C, et al. Effect of intermediate-frequency repetitive transcranial magnetic stimulation on recovery following traumatic brain injury in rats. BioMed Res Int. 2017. https://doi.org/10.1155/2017/4540291.

    Article  PubMed  PubMed Central  Google Scholar 

  158. Qian FF, He YH, Du XH, Lu HX, He RH, Fan JZ. Repetitive transcranial magnetic stimulation promotes neurological functional recovery in rats with traumatic brain injury by upregulating synaptic plasticity-related proteins. Neural Regen Res. 2023;18(2):368–74.

    Article  PubMed  Google Scholar 

  159. Yoon KJ, Lee YT, Chae SW, Park CR, Kim DY. Effects of anodal transcranial direct current stimulation (tDCS) on behavioral and spatial memory during the early stage of traumatic brain injury in the rats. J Neurol Sci. 2016;362:314–20.

    Article  PubMed  Google Scholar 

  160. Kim HJ, Han SJ. Anodal transcranial direct current stimulation provokes neuroplasticity in repetitive mild traumatic brain injury in rats. Neural Plast. 2017. https://doi.org/10.1155/2017/1372946.

    Article  PubMed  PubMed Central  Google Scholar 

  161. Bragina OA, Lara DA, Nemoto EM, Shuttleworth CW, Semyachkina-Glushkovskaya OV, Bragin DE. Increases in microvascular perfusion and tissue oxygenation via vasodilatation after anodal transcranial direct current stimulation in the healthy and traumatized mouse brain. In: Thews O, LaManna JC, Harrison DK, editors. Advances in experimental medicine and biology. Cham: Springer, New York LLC; 2018. p. 27–31.

    Google Scholar 

  162. Yu KP, Yoon YS, Lee JG, Oh JS, Lee JS, Seog T, et al. Effects of electric cortical stimulation (ECS) and transcranial direct current stimulation (tDCS) on rats with a traumatic brain injury. Ann Rehabil Med. 2018;42(4):502–13.

    Article  PubMed  PubMed Central  Google Scholar 

  163. Park G, Suh JH, Han SJ. Transcranial direct current stimulation for balance and gait in repetitive mild traumatic brain injury in rats. BMC Neurosci. 2021;22(1):26.

    Article  PubMed  PubMed Central  Google Scholar 

  164. Bragina OA, Semyachkina-Glushkovskaya OV, Nemoto EM, Bragin DE. Anodal transcranial direct current stimulation improves impaired cerebrovascular reactivity in traumatized mouse brain. In: Ryu P-D, LaManna JC, Harrison DK, Lee S-S, editors. Advances in experimental medicine and biology. Springer: Cham; 2020. p. 47–53.

    Google Scholar 

  165. Martens KM, Pechacek KM, Modrak CG, Milleson VJ, Zhu B, Vonder HC. Cathodal transcranial direct-current stimulation selectively decreases impulsivity after traumatic brain injury in rats. J Neurotrauma. 2019;36(19):2827–30.

    Article  PubMed  PubMed Central  Google Scholar 

  166. Lee DJ, Gurkoff GG, Izadi A, Berman RF, Ekstrom AD, Muizelaar JP, et al. Medial septal nucleus theta frequency deep brain stimulation improves spatial working memory after traumatic brain injury. J Neurotrauma. 2013;30(2):131–9.

    Article  PubMed  Google Scholar 

  167. Gonzalez MMC, Blaya MO, Alonso OF, Bramlett HM, Hentall ID. Midbrain raphe stimulation improves behavioral and anatomical recovery from fluid-percussion brain injury. J Neurotrauma. 2013;30(2):119–30.

    Article  Google Scholar 

  168. Lee DJ, Gurkoff GG, Izadi A, Seidl SE, Echeverri A, Melnik M, et al. Septohippocampal neuromodulation improves cognition after traumatic brain injury. J Neurotrauma. 2015;32(22):1822–32.

    Article  PubMed  PubMed Central  Google Scholar 

  169. Aronson JP, Katnani HA, Huguenard A, Mulvaney G, Bader ER, Yang JC, et al. Phasic stimulation in the nucleus accumbens enhances learning after traumatic brain injury. Cereb Cortex Commun. 2022;3(2):tgac016.

    Article  PubMed  PubMed Central  Google Scholar 

  170. Chan HH, Hogue O, Mathews ND, Hunter JG, Kundalia R, Hermann JK, et al. Deep cerebellar stimulation enhances cognitive recovery after prefrontal traumatic brain injury in rodent. Exp Neurol. 2022;355: 114136.

    Article  PubMed  Google Scholar 

  171. Chan HH, Wathen CA, Mathews ND, Hogue O, Modic JP, Kundalia R, et al. Lateral cerebellar nucleus stimulation promotes motor recovery and suppresses neuroinflammation in a fluid percussion injury rodent model. Brain Stimul. 2018;11(6):1356–67.

    Article  PubMed  Google Scholar 

  172. Jen E, Lin CW, Hsieh TH, Chiu YC, Lu TC, Chen SC, et al. Prototype deep brain stimulation system with closed-loop control feedback for modulating bladder functions in traumatic brain injured animals. J Med Biol Eng. 2018;38(3):337–49.

    Article  Google Scholar 

  173. Praveen Rajneesh C, Lai CH, Chen SC, Hsieh TH, Chin HY, Peng CW. Improved voiding function by deep brain stimulation in traumatic brain-injured animals with bladder dysfunctions. Int Urol Nephrol. 2019;51(1):41–52.

    Article  PubMed  Google Scholar 

  174. Rajneesh CP, Liou JC, Hsieh TH, Chin HY, Peng CW. Efficacy of deep brain stimulation on the improvement of the bladder functions in traumatic brain injured rats. Brain Sci. 2020;10(11):1–13.

    Google Scholar 

  175. Tabansky I, Quinkert AW, Rahman N, Muller SZ, Lofgren J, Rudling J, et al. Temporally-patterned deep brain stimulation in a mouse model of multiple traumatic brain injury. Behav Brain Res. 2014;273:123–32.

    Article  PubMed  PubMed Central  Google Scholar 

  176. Dong X, Ye W, Tang Y, Wang J, Zhong L, Xiong J, et al. Wakefulness-promoting effects of lateral hypothalamic area-deep brain stimulation in traumatic brain injury-induced comatose rats: upregulation of α1-adrenoceptor subtypes and downregulation of gamma-aminobutyric acid β receptor expression via the orexins pathway. World Neurosurg. 2021;152:e321–31.

    Article  PubMed  Google Scholar 

  177. Smith DC, Modglin AA, Roosevelt RW, Neese SL, Jensen RA, Browning RA, et al. Electrical stimulation of the vagus nerve enhances cognitive and motor recovery following moderate fluid percussion injury in the rat. J Neurotrauma. 2005;22(12):1485–502.

    Article  PubMed  Google Scholar 

  178. Smith DC, Tan AA, Duke A, Neese SL, Clough RW, Browning RA, et al. Recovery of function after vagus nerve stimulation initiated 24 hours after fluid percussion brain injury. J Neurotrauma. 2006;23(10):1549–60.

    Article  PubMed  Google Scholar 

  179. Pruitt DT, Schmid AN, Kim LJ, Abe CM, Trieu JL, Choua C, et al. Vagus nerve stimulation delivered with motor training enhances recovery of function after traumatic brain injury. J Neurotrauma. 2016;33(9):871–9.

    Article  PubMed  PubMed Central  Google Scholar 

  180. Dong XY, Feng Z. Wake-promoting effects of vagus nerve stimulation after traumatic brain injury: upregulation of orexin-A and orexin receptor type 1 expression in the prefrontal cortex. Neural Regen Res. 2018;13(2):244–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  181. Dong X, Papa E, Liu H, Feng Z, Huang F, Liao C. Vagus nerve stimulation causes wake-promotion by affecting neurotransmitters via orexins pathway in traumatic brain injury induced comatose rats. Int J Clin Exp Med 2018;11(5):4742–51.

  182. Clough RW, Neese SL, Sherill LK, Tan AA, Duke A, Roosevelt RW, et al. Cortical edema in moderate fluid percussion brain injury is attenuated by vagus nerve stimulation. Neuroscience. 2007;147(2):286–93.

    Article  CAS  PubMed  Google Scholar 

  183. Zhou L, Lin J, Lin J, Kui G, Zhang J, Yu Y. Neuroprotective effects of vagus nerve stimulation on traumatic brain injury. Neural Regen Res. 2014;9(17):1585–91.

    Article  PubMed  PubMed Central  Google Scholar 

  184. Neese SL, Sherill LK, Tan AA, Roosevelt RW, Browning RA, Smith DC, et al. Vagus nerve stimulation may protect GABAergic neurons following traumatic brain injury in rats: an immunocytochemical study. Brain Res. 2007;1128(1):157–63.

    Article  CAS  PubMed  Google Scholar 

  185. Ping X, Jin X. Transition from initial hypoactivity to hyperactivity in cortical layer V pyramidal neurons after traumatic brain injury in vivo. J Neurotrauma. 2016;33(4):354–61.

    Article  PubMed  PubMed Central  Google Scholar 

  186. Carron SF, Alwis DS, Rajan R. Traumatic brain injury and neuronal functionality changes in sensory cortex. Front Syst Neurosci. 2016;10:47.

    Article  PubMed  PubMed Central  Google Scholar 

  187. Walton D, Spencer DC, Nevitt SJ, Michael BD. Transcranial magnetic stimulation for the treatment of epilepsy. Cochrane Database Syst Rev. 2021. https://doi.org/10.1002/14651858.CD011025.pub3/full.

    Article  PubMed  PubMed Central  Google Scholar 

  188. Huttunen JK, Airaksinen AM, Barba C, Colicchio G, Niskanen JP, Shatillo A, et al. Detection of hyperexcitability by functional magnetic resonance imaging after experimental traumatic brain injury. J Neurotrauma. 2018;35(22):2708–17.

    Article  PubMed  Google Scholar 

  189. Kaloostian P, Robertson C, Gopinath SP, Stippler M, King CC, Qualls C, et al. Outcome prediction within twelve hours after severe traumatic brain injury by quantitative cerebral blood flow. J Neurotrauma. 2012;29(5):727–34.

    Article  PubMed  PubMed Central  Google Scholar 

  190. Li DH, Yang XF. Remote modulation of network excitability during deep brain stimulation for epilepsy. Seizure. 2017;47:42–50.

    Article  PubMed  Google Scholar 

  191. Torres Diaz CV, González-Escamilla G, Ciolac D, Navas García M, Pulido Rivas P, Sola RG, et al. Network substrates of centromedian nucleus deep brain stimulation in generalized pharmacoresistant epilepsy. Neurotherapeutics. 2021;18(3):1665–77.

    Article  PubMed  PubMed Central  Google Scholar 

  192. Khambhati AN, Shafi A, Rao VR, Chang EF. Long-term brain network reorganization predicts responsive neurostimulation outcomes for focal epilepsy. Sci Transl Med. 2021;13(608):eabf6588.

    Article  PubMed  Google Scholar 

  193. Ponsford J, Willmott C, Rothwell A, Cameron P, Kelly AM, Nelms R, et al. Impact of early intervention on outcome following mild head injury in adults. J Neurol Neurosurg Psychiatry. 2002;73(3):330–2.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  194. Andelic N, Bautz-Holter E, Ronning P, Olafsen K, Sigurdardottir S, Schanke AK, et al. Does an early onset and continuous chain of rehabilitation improve the long-term functional outcome of patients with severe traumatic brain injury? J Neurotrauma. 2012;29(1):66–74.

    Article  PubMed  Google Scholar 

  195. Oberholzer M, Müri RM. Neurorehabilitation of traumatic brain injury (TBI): a clinical review. Med Sci. 2019;7(3):47.

    Google Scholar 

  196. Laing J, Gabbe B, Chen Z, Perucca P, Kwan P, O’Brien TJ. Risk factors and prognosis of early posttraumatic seizures in moderate to severe traumatic brain injury. JAMA Neurol. 2022;79(4):334–41.

    Article  PubMed  PubMed Central  Google Scholar 

  197. Hakon J, Moghiseh M, Poulsen I, Øland CML, Hansen CP, Sabers A. Transcutaneous vagus nerve stimulation in patients with severe traumatic brain injury: a feasibility trial. Neuromodul Technol Neural Interface. 2020;23(6):859–64.

    Article  Google Scholar 

  198. Zhao YX, He W, Jing XH, Liu JL, Rong PJ, Ben H, et al. Transcutaneous auricular vagus nerve stimulation protects endotoxemic rat from lipopolysaccharide-induced inflammation. Evid Based Complement Alternat Med. 2012;2012: e627023.

    Article  Google Scholar 

  199. Imazawa W, Nakamura H, Yagi M, Morishita K, Otomo Y, Ueno A. Measurement of vagus nerve response to transcutaneous electrical ear canal stimulation in anesthetized rat. In: 2020 42nd annual international conference of the IEEE Engineering in Medicine & Biology Society (EMBC). 2020. p. 5216–9.

  200. Long L, Zang Q, Jia G, Fan M, Zhang L, Qi Y, et al. Transcutaneous auricular vagus nerve stimulation promotes white matter repair and improves dysphagia symptoms in cerebral ischemia model rats. Front Behav Neurosci. 2022. https://doi.org/10.3389/fnbeh.2022.811419.

    Article  PubMed  PubMed Central  Google Scholar 

  201. He W, Jing XH, Zhu B, Zhu XL, Li L, Bai WZ, et al. The auriculo-vagal afferent pathway and its role in seizure suppression in rats. BMC Neurosci. 2013;14(1):85.

    Article  PubMed  PubMed Central  Google Scholar 

  202. Johnson RL, Wilson CG. A review of vagus nerve stimulation as a therapeutic intervention. J Inflamm Res. 2018;11:203–13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  203. Silverman J, Muir WW. A review of laboratory animal anesthesia with chloral hydrate and chloralose. Lab Anim Sci. 1993;43(3):210–6.

    CAS  PubMed  Google Scholar 

  204. Ellenbroek B, Youn J. Rodent models in neuroscience research: is it a rat race? Dis Model Mech. 2016;9(10):1079–87.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  205. Tan AA, Quigley A, Smith DC, Hoane MR. Strain differences in response to traumatic brain injury in Long-Evans compared to Sprague–Dawley rats. J Neurotrauma. 2009;26(4):539–48.

    Article  PubMed  PubMed Central  Google Scholar 

  206. Xiong Y, Mahmood A, Chopp M. Animal models of traumatic brain injury. Nat Rev Neurosci. 2013;14(2):128–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  207. Serpa RO, Ferguson L, Larson C, Bailard J, Cooke S, Greco T, et al. Pathophysiology of pediatric traumatic brain injury. Front Neurol. 2021. https://doi.org/10.3389/fneur.2021.696510.

    Article  PubMed  PubMed Central  Google Scholar 

  208. Gupte RP, Brooks WM, Vukas RR, Pierce JD, Harris JL. Sex differences in traumatic brain injury: what we know and what we should know. J Neurotrauma. 2019;36(22):3063–91.

    Article  PubMed  PubMed Central  Google Scholar 

  209. Späni CB, Braun DJ, Van Eldik LJ. Sex-related responses after traumatic brain injury: considerations for preclinical modeling. Front Neuroendocrinol. 2018;50:52–66.

    Article  PubMed  PubMed Central  Google Scholar 

  210. Valera EM, Joseph ALC, Snedaker K, Breiding MJ, Robertson CL, Colantonio A, et al. Understanding traumatic brain injury in females: a state-of-the-art summary and future directions. J Head Trauma Rehabil. 2021;36(1):E1.

    Article  PubMed  PubMed Central  Google Scholar 

  211. Laskowski RA, Creed JA, Raghupathi R. Pathophysiology of mild TBI: implications for altered signaling pathways. In: Kobeissy FH, editor. Brain neurotrauma: molecular, neuropsychological, and rehabilitation aspects (frontiers in neuroengineering). Boca Raton: CRC Press/Taylor & Francis; 2015.

    Google Scholar 

  212. da Silva TH, Massetti T, da Silva TD, da Silva Paiva L, Papa DCR, de Mello Monteiro CB, et al. Influence of severity of traumatic brain injury at hospital admission on clinical outcomes. Fisioter E Pesqui. 2018;25:3–8.

    Article  Google Scholar 

  213. Andriessen TMJC, Horn J, Franschman G, van der Naalt J, Haitsma I, Jacobs B, et al. Epidemiology, severity classification, and outcome of moderate and severe traumatic brain injury: a prospective multicenter study. J Neurotrauma. 2011;28(10):2019–31.

    Article  PubMed  Google Scholar 

  214. Maasdorp SD, Swanepoel C, Gunter L. Outcomes of severe traumatic brain injury at time of discharge from tertiary academic hospitals in Bloemfontein. Afr J Thorac Crit Care Med. 2020;26(2):32–5.

    Article  Google Scholar 

  215. Marklund N, Bellander BM, Godbolt AK, Levin H, McCrory P, Thelin EP. Treatments and rehabilitation in the acute and chronic state of traumatic brain injury. J Intern Med. 2019;285(6):608–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  216. Khan F, Baguley IJ, Cameron ID. 4: rehabilitation after traumatic brain injury. Med J Aust. 2003;178(6):290–5.

    Article  PubMed  Google Scholar 

  217. Iaccarino MA, Bhatnagar S, Zafonte R. Chapter 26—rehabilitation after traumatic brain injury. In: Grafman J, Salazar AM, editors. Handbook of clinical neurology (traumatic brain injury, part I), vol. 127. Amsterdam: Elsevier; 2015. p. 411–22.

    Google Scholar 

  218. Huashuang Z, Yang L, Chensheng H, Jing X, Bo C, Dongming Z, et al. Prevalence of adverse effects associated with transcranial magnetic stimulation for autism spectrum disorder: a systematic review and meta-analysis. Front Psychiatry. 2022;13: 875591.

    Article  PubMed  PubMed Central  Google Scholar 

  219. Brunoni AR, Amadera J, Berbel B, Volz MS, Rizzerio BG, Fregni F. A systematic review on reporting and assessment of adverse effects associated with transcranial direct current stimulation. Int J Neuropsychopharmacol. 2011;14(8):1133–45.

    Article  PubMed  Google Scholar 

  220. Ben-Menachem E. Vagus nerve stimulation, side effects, and long-term safety. J Clin Neurophysiol. 2001;18(5):415–8.

    Article  CAS  PubMed  Google Scholar 

  221. Zarzycki MZ, Domitrz I. Stimulation-induced side effects after deep brain stimulation—a systematic review. Acta Neuropsychiatr. 2020;32(2):57–64.

    Article  PubMed  Google Scholar 

  222. Papageorgiou PN, Deschner J, Papageorgiou SN. Effectiveness and adverse effects of deep brain stimulation: umbrella review of meta-analyses. J Neurol Surg A. 2017;78(2):180–90.

    Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

Open access funding provided by Austrian Science Fund (FWF). This research was funded in whole by the Austrian Science Fund (FWF Project No. ZK 17 Zukunftskolleg to MÜ, KK, SS, and TR).

Author information

Authors and Affiliations

Authors

Contributions

DZ: Conceptualization, Methodology, Validation, Investigation, Data Curation, Writing—Original Draft, Writing—Review & Editing, Visualization. MN: Conceptualization, Validation, Investigation, Writing—Original Draft, Writing—Review & Editing. MÜ: Conceptualization, Validation, Investigation, Writing—Original Draft, Writing—Review & Editing, Project administration, Funding acquisition. SS, KK: Writing Review & Editing, Funding acquisition. US, RS and CB: Writing—Review & Editing. TR: Conceptualization, Validation, Investigation, Writing—Original Draft, Writing—Review & Editing, Supervision, Project administration, Funding acquisition. All authors read and approved the final manuscript.

Corresponding author

Correspondence to T. Rienmüller.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ziesel, D., Nowakowska, M., Scheruebel, S. et al. Electrical stimulation methods and protocols for the treatment of traumatic brain injury: a critical review of preclinical research. J NeuroEngineering Rehabil 20, 51 (2023). https://doi.org/10.1186/s12984-023-01159-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12984-023-01159-y

Keywords