Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2013 Jul 1.
Published in final edited form as: Brain Stimul. 2011 Apr 1;5(3):175–195. doi: 10.1016/j.brs.2011.03.002

Clinical Research with Transcranial Direct Current Stimulation (tDCS): Challenges and Future Directions

Andre Russowsky Brunoni 1, Michael A Nitsche 2, Nadia Bolognini 3, Marom Bikson 4, Tim Wagner 5, Lotfi Merabet 6, Dylan J Edwards 7, Antoni Valero-Cabre 8, Alexander Rotenberg 9, Alvaro Pascual-Leone 10, Roberta Ferrucci 11, Alberto Priori 12, Paulo Boggio 13, Felipe Fregni 14
PMCID: PMC3270156  NIHMSID: NIHMS283820  PMID: 22037126

Abstract

Background

Transcranial direct current stimulation (tDCS) is a neuromodulatory technique that delivers low-intensity, direct current to cortical areas facilitating or inhibiting spontaneous neuronal activity. In the past ten years, tDCS physiological mechanisms of action have been intensively investigated giving support for the investigation of its applications in clinical neuropsychiatry and rehabilitation. However, new methodological, ethical, and regulatory issues emerge when translating the findings of preclinical and phase I studies into phase II and III clinical studies. The aim of this comprehensive review is to discuss the key challenges of this process and possible methods to address them.

Methods

We convened a workgroup of researchers in the field to review, discuss and provide updates and key challenges of neuromodulation use for clinical research.

Main Findings/Discussion

We reviewed several basic and clinical studies in the field and identified potential limitations, taking into account the particularities of the technique. We review and discuss the findings into four topics: (i) mechanisms of action of tDCS, parameters of use and computer-based human brain modeling investigating electric current fields and magnitude induced by tDCS; (ii) methodological aspects related to the clinical research of tDCS as divided according to study phase (i.e., preclinical, phase I, phase II and phase III studies); (iii) ethical and regulatory concerns; (iv) future directions regarding novel approaches, novel devices, and future studies involving tDCS. Finally, we propose some alternative methods to facilitate clinical research on tDCS.

Keywords: Transcranial direct current stimulation, brain stimulation, clinical research, physical medicine, neuropsychiatry, medical devices

1. Introduction

The effects of uncontrolled electrical stimulation on the brain have been reported since the distant past. Scribonius Largus (the physician of the Roman Emperor Claudius), described how placing a live torpedo fish over the scalp to deliver a strong electric current could relieve a headache (9). Galen of Pergamum, the great medical researcher of the ancients, and Pliny the Elder also described similar findings (11). In the 11th Century, Ibn-Sidah, a Muslim physician, suggested using a live electric catfish for the treatment of epilepsy (11).

With the introduction of the electric battery in the 18th Century, it became possible to evaluate the effect of direct transcranial stimulation systematically. Individuals such as Walsh (1773), Galvani (1791, 1797) and Volta (1792) all recognized that electrical stimulation of varying duration could evoke different physiological effects (14). In fact, one of the first systematic reports of clinical applications of galvanic currents date back to this period, when Giovanni Aldini (Galvani’s nephew) and others used transcranial electrical stimulation to treat melancholia (19, 20). Over the past two centuries, many other researchers (see Zago et al. (14) for further references) used galvanic current for the treatment of mental disorders with varying results. In more recent history, the use of electroconvulsive therapy and psychopharmacological drugs and lack of reliable neurophysiological markers have obscured direct current stimulation of the central nervous system as a therapeutic and research tool particularly in the field of psychiatry. Nonetheless, galvanic current has been used without interruption for the treatment of musculoskeletal disorders and peripheral pain.

In fact, a reappraisal of transcranial direct current stimulation (tDCS) as a form of noninvasive brain stimulation took place at the turn of this century. The seminal studies of Priori and colleagues (27) followed by Nitsche and Paulus (28) demonstrated that weak, direct electric currents could be delivered effectively transcranially as to induce bidirectional, polarity-dependent changes in cortical. Specifically, anodal direct current stimulation was shown to increase cortical excitability, whereas cathodal stimulation decreased it. In addition, animal and human studies have provided insight regarding the mechanisms underlying tDCS effects on neuroplasticity (2932) and current distribution according to the brain area being stimulated (3336). In addition, several studies showed that tDCS could induce specific changes in neuropsychological, psychophysiological and motor activity as a function of targeted brain areas (3, 3739). Moreover, certain appealing characteristics of tDCS (such as the fact that it is non-invasive and has mostly well-tolerated, transient and mild adverse effects) have sparked an increase in clinical studies particularly for neuropsychiatric disorders such as major depressive disorder, chronic and acute pain, stroke rehabilitation, drug addiction and other neurological and psychiatric conditions (4042). Although reported effects have been heterogeneous and warrant further clinical studies, studies have been generally promising.

As the field of noninvasive brain stimulation moves towards more clinical applications, there are new issues that emerge. One is methodological; how to study tDCS in neuropsychiatry that historically has been heavily pharmacotherapy-based (43). Specifically, what are the optimal approaches regarding study design (e.g. two-arm, three-arm versus factorial), study methodology (blinding, use of placebo, concomitant use of drugs) sample requirements (i.e. sample size, eligibility criteria, sample recruitment), interventions (e.g. electrode positioning, dosage, duration, and also comparison against pharmacotherapy), outcomes (e.g. clinical vs. surrogate outcomes) and safety. Another issue is ethical; who should apply tDCS in clinical settings (e.g. physicians, neuropsychologists, specialized staff); the tolerable amount of risk for inducing maladaptive, long-term neuroplasticity, and whether tDCS could be used for enhancing neuropsychological performance in healthy subjects; finally, regulatory issues also need to be discussed. In contrast to transcranial magnetic stimulation (TMS), which is delivered through a sophisticated device, transcranial direct current stimulation can be administered with devices already manufactured and used in pain and cosmetic medicine. These devices deliver direct current to the joints and/or the skin. Also, contrary to TMS, these devices are affordable and readily accessible and can be purchased by non-trained individuals, including patients.

The last question is why conducting clinical research on tDCS. Among others, we can identify three main reasons: 1) there is a theoretical clinical basis for tDCS as a substitutive treatment for pharmacotherapy, such as patients with poor drug tolerability or those with adverse pharmacological interactions (e.g. elderly people who use several drugs). For instance, one group that would potentially benefit from further investigation of tDCS safety is pregnant women with unipolar depression, as there is a lack of acceptable pharmacological alternatives for this condition (44). 2) Using tDCS as an augmentative treatment - e.g., tDCS and restraint therapy for stroke (45), or tDCS and pharmacotherapy for chronic pain or major depression. Again, side effects and non-invasiveness make tDCS an appealing strategy to boost the effects of other treatments in addition to its neurophysiological effects on membrane resting threshold that likely underlie its synergistic effects. 3) tDCS is inexpensive; being therefore attractive to areas lacking in resources. If proven effective, tDCS will be an interesting option for developing countries.

The purpose of this review is to assess the current stage of tDCS development and identify its potential limitations in current clinical studies as to provide a comprehensive framework for designing future clinical trials. This review is divided in four parts. The first part reviews the mechanisms of action of tDCS, parameters of use and computer-based human brain modeling investigating electric current fields and magnitude induced by tDCS. Given the conciseness of this section, the reader is invited to consult more recent reviews focusing exclusively on the mechanisms of action and technical development (see (46) and (47)). The second section covers methodological aspects related to the clinical research application of tDCS. This section is divided according to study phase (i.e., preclinical, phase I, phase II and phase III studies). The third section focuses on ethical and regulatory concerns. The last section concludes with a presentation of what are expected in the near future regarding novel approaches, novel devices, and future studies involving tDCS.

2. The electrophysiology of transcranial direct current stimulation

2.1 Mechanisms of action

TDCS differs from other non-invasive brain stimulation techniques such as transcranial electrical stimulation (TES) and TMS. TDCS does not induce neuronal firing by suprathreshold neuronal membrane depolarization but rather modulates spontaneous neuronal network activity (47, 48). At the neuronal level, the primary mechanism of action is a tDCS polarity-dependent shift (polarization) of resting membrane potential. While anodal DC stimulation generally enhances cortical activity and excitability, cathodal DC stimulation has opposite effects (28, 49, 50). Animal studies have shown that changes in excitability are reflected in both spontaneous firing rates (51, 52); and responsiveness to afferent synaptic inputs (53, 54). It is this primary polarization mechanism that underlies the acute effects of low-intensity DC currents on cortical excitability in humans (27).

However, tDCS elicits after-effects lasting for up to one hour (30, 55). Therefore, its mechanisms of action cannot be solely attributed to changes of the electrical neuronal membrane potential. In fact, further research showed that tDCS also modifies the synaptic microenvironment, for instance, by modifying synaptic strength NMDA receptor-dependently or altering GABAergic activity (5658). TDCS also interferes with brain excitability through modulation of intracortical and corticospinal neurons (31, 59). The effects of tDCS might be similar to those observed in long-term potentiation (LTP), as shown by one recent animal study that applied anodal motor cortex stimulation and showed a lasting increase in postsynaptic excitatory potentials (29). Experiments with peripheral nerve (59) and spinal cord (60) stimulation showed that DC effects are also non-synaptic, possibly involving transient changes in the density of protein channels localized below the stimulating electrode.

Given that a constant electric field displaces all polar molecules and most of the neurotransmitters and receptors in the brain have electrical properties, tDCS might also influence neuronal function by inducing prolonged neurochemical changes (58, 61). For instance, magnetic resonance spectroscopy showed that after anodal tDCS brain myoinositol significantly increased whereas n-acetyl-aspartate failed to change (61).

In addition to the “direct” tDCS effects described above, “indirect” effects are also observed. This is seen in connectivity-driven alterations of distant cortical and subcortical areas (62, 63). Interestingly, tDCS modulates not only single neuron activity and evoked neuronal activity, but also spontaneous neuronal oscillations. Ardolino et al. (59) found that below the cathodal electrode, the slow EEG activity in the theta and delta band increases. Animal and modeling studies suggest that a network of tightly coupled active neurons (e.g. oscillations) may be more sensitivity to applied weak current than neurons in isolation (6466).

Although most early tDCS studies have been performed in the motor cortex, it should be noticed that tDCS does not only induce long-lasting alterations of motor-evoked potentials, but also affects somatosensory and visual-evoked potentials. This activity is dependent on the area stimulated (6769). Ferrucci et al. (8) and Galea et al. (70) provided evidence that tDCS can influence the human cerebellum. Cogiamanian et al. (60) and Winkler et al. (71) demonstrated that transcutaneous DC stimulation modulates conduction along the spinal cord and the segmental reflex pathways.

An important aspect when discussing the mechanisms of tDCS is the magnitude and location of the current induced in cortical tissues. Several modeling studies have been conducted to address this issue and will be discussed in a later section.

Finally, constant electrical fields influence several different tissues (vessels, connective tissue etc.) and pathophysiological mechanisms (inflammation, cell migration, vascular motility); in addition, their effects are observed on multiple cellular structures (cytoskeleton, mithocondria, membrane). With that said, tDCS may also influence non-neuronal components of the central nervous system. Support for this theory is observed below anodal tDCS electrode as it can induce prolonged brain vasodilatation (72).

In conclusion, the mechanisms of action of DCS remain to be completely elucidated, an issue that can have important repercussions for future clinical applications. These mechanisms likely involve different synaptic and non-synaptic effects on neurons and effects on non-neuronal cells and tissues within the central nervous system.

2.2 Pharmacological investigation of transcranial direct current stimulation

In tDCS research, pharmacological studies use diverse drugs to block and/or enhance the activity of neurotransmitters and its receptors to observe how and whether tDCS-induced cortical excitability is modified. Therefore, such studies aim to enhance our knowledge about the mechanisms of action of tDCS with regard to neuromodulation and neuroplasticity.

Evidence suggests that blocking voltage-gated sodium and calcium channels decreases the excitability-enhancing effect of anodal tDCS. In contrast, cathodal tDCS-generated excitability reductions are not affected (56, 57). These findings are in line with the assumption that tDCS induces shifts in membrane resting threshold of cortical neurons.

Regarding neurotransmitters, it has been shown that NMDA-glutamatergic receptors are involved in inhibitory and facilitatory plasticity induced by tDCS. Blocking NMDA receptors abolishes the after-effects of stimulation, while enhancement of NMDA receptor efficacy by d-cycloserine enhances selectively facilitatory plasticity (30, 73). In contrast, GABAergic modulation with lorazepam results in a delayed then enhanced and prolonged anodal tDCS-induced excitability elevation (74) (Table 1).

Table 1.

Pharmacological agents that interact with transcranial direct current stimulation effects on cortical excitability.

Drug Class Effect
Amine metabolism
Citalopram SERT blocker Enhancement of the duration of facilitatory anodal effects; Facilitation of cathodal tDCS effects (78)
Amphetamine NET/DAT competitive inhibitor Enhancment of the duration of facilitatory anodal effects (74).
L-Dopa Dopamine precursor For anodal: excitability turns into inhibition; For cathodal: effects are enhanced (33)
Sulpiride D2-receptor blocker Abolishment of tDCS-induced plasticity (149)
Pergolide dopamine agonist agent Enhancement of the duration of cathodal tDCS effects (149, 150)
Amino acid metabolism
Lorazepam GABA allosteric modulator Anodal effects are delayed, but then enhanced and prolonged. (104)
Rivastigmine Cholinesterase inhibitor Abolishment of anodal tDCS effects; stabilization of cathodal tDCS effects (151)
Dextromethorpan NMDA antagonist agent Abolishment of the after-effects of anodal and cathodal tDCS (56, 57).
D-cycloserine NMDA agonist agent Enhancement of the duration of anodal effects; no effects during cathodal stimulation (73).
Voltage-sensitive channel blockers
Carbamazepine Voltage-sensitive sodium channel blocker Abolishment of the depolarizing effects of anodal tDCS. (56, 57)
Flunarizine Voltage-sensitive calcium channel blocker Similar effects of Carbamazepine.

tDCS= transcranial direct current stimulation; NET = norepinephrine transporter; DAT = dopamine transporter; GABA = gamma-aminobutyric acid; NMDA = n-methyl-d-aspartic acid; SERT= serotonin transporter.

Regarding the monoaminergic neurotransmitters, amphetamines (that increase monoaminergic activity) seem to enhance tDCS-induced facilitatory plasticity (75). For the dopaminergic system, tDCS-generated plasticity is modulated in a complex dosage- and subreceptor-dependent manner. Application of the dopamine precursor l-dopa converts facilitatory plasticity into inhibition, and prolongs inhibitory plasticity (76), while blocking D2 receptors seems to abolish tDCS-induced plasticity (77), D2 agonists, applied at high or low dosages, decrease plasticity. Furthermore, plasticity is restituted by medium dosage D2 agonists (78). Interestingly, the acetylcholine reuptake-inhibitor rivastigmine affects tDCS-induced plasticity in a similar fashion as l-dopa (32). For the serotoninergic system, the 5-HT reuptake-inhibitor citalopram enhances facilitatory plasticity and also converts inhibitory plasticity into facilitation (79).

From a clinical point of view, these results show that pharmacotherapy and tDCS interact, which might be an issue when studying clinical samples receiving both interventions. In fact, the complex non-linear interaction makes it difficult to foresee the specific effects of pathophysiological alterations or drug application on the amount and direction of tDCS-induced plasticity; thus demanding further empirical research on this topic.

2.2 Parameters of stimulation

TDCS parameters can vary widely and several factors need to be defined. These factors include electrode size and positioning, intensity, duration of stimulation, number of sessions per day, and interval between sessions. By varying these parameters, different amounts of electric current can be delivered, thus inducing diverse physiological and adverse effects. Another potential concern is that tDCS devices are not worldwide standardized. These devices can be easily constructed using standard equipment and technology in engineering laboratories of colleges and universities. In fact, more than a dozen different tDCS devices can be found throughout neuromodulation laboratories worldwide.

2.2.1 Electrode positioning

Although tDCS electrical fields are relatively non-focal, electrode positioning is critical. For instance, a previous study showed that changing the electrode reference from DLPFC to M1 eliminated tDCS effects on working memory (3). Other studies have shown that phosphene-thresholds are modulated only during occipital (visual cortex) DC stimulation and not other areas (69, 80). Likewise, a tDCS trial for major depression showed that only DLPFC stimulation (and not occipital stimulation) ameliorated symptoms (2). Although current evidence suggests site-dependent effects, other issues have yet to be explored – for instance, one open question is whether and how brain stimulation in one area influences adjacent and more distant areas.

TDCS studies usually use one anode and one cathode electrode placed over the scalp to modulate a particular area of the CNS. Electrode positioning is usually determined according to the International EEG 10–20 System. Given the focality of tDCS, this appears appropriate. For instance, studies exploring the motor cortex place electrodes over C3 or C4; for the visual system, electrodes are typically placed over O1 or O2 (for a review of tDCS studies exploring different brain areas see Utz et al. (81)).

Here, some terms used to describe tDCS montages should be discussed: when one electrode is placed bellow the neck, the entire montage is usually described as “unipolar”. On the other hand, montages with two electrodes on the head are termed usually “bipolar”. However, this nomenclature might be inaccurate as technically the DC stimulation is always generated via two poles (electrodes) generating an electric dipole between the electrodes. Therefore, an alternative nomenclature of “mono-cephalic” and “bi-cephalic” is proposed to differentiate between “unipolar” and “bipolar” setups, respectively. Researchers in the field also use the terms “reference” and “stimulating” electrode to refer to the “neutral” and “active” electrode, respectively. However, the term “reference” electrode may also be problematic, especially for bicephalic montages because the “reference” electrode is not physiologically inert and can contribute to activity modulation as well. This could be a potential confounder depending on the main study question. Nonetheless, researchers use these terms to highlight that (in their study) they are under the assumption that in their particular montage one electrode is being explored as the “stimulating” while the other is the “reference”.

On the other hand, having the possibility to increase and decrease activity in different brain areas simultaneously may be advantageous. For instance, this could be useful in conditions involving an imbalanced interhemispheric activity (i.e., in stroke) (82). In scenarios whether the reference electrode poses a confounding effect, an extra-cephalic reference electrode could theoretically aid in avoiding this issue. However, this might increase the risk of shunting the electric current through the skin (which would then not reach brain tissue) or displacing the current. Ultimately, the choice of montage will be application specific; for example a recent study comparing different tDCS setups showed that, although bicephalic setups were effective, the monocephalic setup was no different than sham stimulation (83). Finally, in a monocephalic setup, using very high currents there is the potential risk of influencing brain stem activity, including respiratory control (note that this risk is theoretical and was only observed in one historical report (84) Nevertheless, in choosing the extracephalic position, the researcher must be confident that a significant electric field will be induced on the target brain area.

Moreover, since current flow direction/electrical field orientation relative to neuronal orientation might determine the effects of tDCS (28), it might be that the effects of an extracephalic electrode differs relevantly from that of a bipolar electrode arrangement. Alternatively, enhancing the size of one electrode, thus reducing current density, might enable functional monocephalic stimulation also with a bicephalic electrode montage (77).

Direct current stimulation can also be delivered over non-cortical brain areas. Ferrucci et al. (8) stimulated the cerebellum showing changes in performance in a cognitive task for working memory. Galea et al. (70) explored the inhibitory effects of the cerebellum on motor-evoked potentials (MEP) triggered by TMS over the motor cortex. This revealed that tDCS could modify MEPs in a polarity-specific manner. In addition, Cogiamanian et al. (60) observed that cathodal transcutaneous DC over the thoracic spinal cord suppressed tibial somatosensory-evoked potentials. Furthermore, Winkler et al. (71) observed that transcutaneous DC stimulation over the spinal cord modulates the post activation depression of the H-reflex. Preliminary data indicates spinal DC stimulation also influences nociception (85) suggesting that the spinal cord as a target for transcutaneous DC stimulation. Challenges for stimulation in this area must be considered such as location of induced electrical fields.

2.2 Modeling transcranial direct current stimulation

During tDCS, current is generated across the brain; different montages result in distinct current flow through the brain and thus the ability to adjust montage allows customization and optimization of tDCS for specific applications (see above). Though tDCS montage design often follow basic assumptions (e.g. “increased/decreased excitability under the anode/cathode”), computational models of brain current flow during tDCS (so called “forward” models) provide more accurate insight into detailed current flow patterns, and in some cases show that the basic assumptions are not valid. When interpreting the results of such simulations, it is important to recognize that the intensity of current flow in any specific brain region does not translate in any simple linear manner to the degree of brain modulation. However, it seems reasonable to predict that regions with more current flow are more likely to be affected by stimulation while regions with little or no current flow will be spared the direct effects of stimulation.

Computational models of tDCS range in complexity from concentric sphere models to individualized high-resolution models based an individual’s structural MRI. The appropriate level of detail depends on the available computational resources and the clinical question being asked (see technical note below). Regardless of complexity, all models share the primary outcome of correctly predicting brain current flow during transcranial stimulation to guide clinical practice in a meaningful manner.

Most clinical studies use tDCS devices that apply direct electric currents via a constant current source, but even within this space there are infinite variations of dosage and montage that can be leveraged, with the help of models, to optimize outcomes. The current is sent through patch electrodes (surface areas typical range from 25–35 cm2 but can vary by an order of magnitude) attached to the scalp surface. Total current injected ranges in magnitude are typically from 0.5 to 2 mA. Steps taken to improve tDCS specificity (including the use of larger “return” sponges and extra-cephalic electrodes) have been proposed but more analysis is required to determine the role of electrode-montage in neuro-modulation and targeting. Modeling approaches are instrumental toward this goal. For example, modeling studies have recently predicted a profound role of the “return” electrode position in modulating overall current flow including under the “active” (or “stimulating”) electrode (86). Specifically, for a fixed active electrode position on the head, changing the position of the return electrode (including cephalic and extracephalic positions) influences current flow through the presumed target region directly under the active electrode. Therefore, in addition to considering the role of scalp shunting and action on deep brain structures (see above) when determining electrode distance, the complete design of electrode montage may subtly modulate cortical current flow (87). Again, computer modeling can provide valuable insight into this process.

Recent modeling studies suggest that individual anatomical differences may affect current flow through the cortex. In comparison to TMS, which uses motor-evoked potentials to index its potency, there is no similar rationale for titrating tDCS dosage. A related issue is the modification of tDCS dose montages for individuals with skull defects or stroke-related lesions. Such individuals may be candidates for tDCS therapy but defects/lesions are expected to distort current flow. For example any defect/injury filled with cerebrospinal fluid (CSF), including those related to stroke of traumatic brain injury, is expected to preferentially “shunt” current flow (36). Ideally, tDCS would be adjusted in a patient-specific (defect/lesion specific) manner to take advantage of such distortions in guiding current flow to targeted regions, while simultaneously avoiding any safety concerns (such as current hot spots).

Evidence from modeling studies suggests that for typical tDCS significant amounts of current can reach broad cortical areas especially between and under the electrode surface (33, 34). Modeling studies also show that electrode montage is critical to the amount of current shunted through the skin.

Electrode montage is critically associated to the amount of current being shunted through the skin, how much is delivered to the brain, and to what targets. The overall theme emerging from modeling efforts is that despite clinical success in applying simplifying rules in dose design, all the details and aspects of electrode montage design combine to influence current flow such that these simplifying rules are applicable but only within a limited parameter range. For example, average current density (total current/electrode area) at the “active” electrode may be a useful metric to normalize specific neurophysiologic outcomes (e.g. TMS evoked MEPs), there is no universal relationship between current density and brain modulation when one considers the full spectrum of possible electrode montages (34, 88).

Recent modeling data taking into consideration gyri and sulci geometry have shown that electric current can concentrate on the edge of gyri (89). Therefore, the effects might not be homogeneous throughout the stimulated area. Increased appreciation of the complexity of current flow through the head (reflecting the complexity of neuro-anatomy), reinforces the utility of applying computational models to assist in tDCS dose design (90) rather then simply relying on some heuristic rules (e.g. “increased excitability under the anode”).

In addition to predicting brain current flow, modeling studies also provide insight into electrode design by predicting current flow patterns through the skin. Modeling studies has reinforced that current is not passed uniformly through the skin but rather tends to concentrate near electrode edges or skin inhomogeneities (34). Electrode design can be simple saline-soaked cotton or sponge pads or specifically designed patches with unique shapes and materials to maximize stimulation magnitude and focality. Modeling confirms that decreasing the salinity of the pads reduces peak current concentration at the edges (even as the total current applied and average current density is fixed) (91).

In summary, modeling studies are expected to play a critical role in the development of next-generation tDCS technologies and approaches. Notably, tDCS devices have not drastically changed since the time when the battery was first discovered. Thus, conventional technology has certain limitations. These include focality (area stimulated), depth of penetration, and targeting-location control. To overcome these and other limitations, technologies using arrays of electrodes (92) such as ‘High Definition’ tDCS (HD-tDCS) (89) and others (e.g. simultaneous EEG monitoring during tDCS as to adjust dosage and parameters) have been recently proposed. Ultimately, as we begin integrating modern technology with transcranial stimulation techniques, clinical control and efficacy will undoubtedly improve.

On a final technical note: Though there has been a recent emphasize to develop increasingly accurate and complex models (89, 90, 93), certain universal technical issues should be considered for high-precision models, beginning with: 1) high-resolution (e.g. 1 mm) anatomical scans so that the entire model work flow should preserve precision. Any finite-element human head model is limited by the precision and accuracy of tissue dimensions (masks) and conductivity values incorporated (inhomogeneity and anisotropy). One hallmark of precision is the cortical surface used in the final finite-element mask solver should represent realistic sulci and gyri; 2) Simultaneously, a priori knowledge of tissue anatomy and factors know known to shape current flow are applied to further refine segmentation. Particularly critical are discontinuities not present in nature that result from limited scan resolution; notably both unnatural perforations in planar tissues (e.g. holes in cerebrospinal fluid where brain contacts skull) and microstructures (e.g. incomplete or voxelized vessels) can produce significant aberrations in predicted current flow. Addition of complexity without proper parameterization can evidently decrease prediction accuracy. An improper balance between these factors can lead to distortions in brain current flow of an order of magnitude or more – uncontrolled additional complexity can in fact induce distortion. We thus emphasize that the most appropriate methodology (ranging from concentric spheres to individualized models) ultimately depends on the clinical question being addressed.

Box 1 – Insight from tDCS studies on cognition.

TDCS has been increasingly used to transiently modify cognitive functions in the healthy human brain. This field presents an exciting opportunity to extend the application of tDCS from a neuroscience research tool to the potential treatment of cognitive impairments. Indeed, the understanding of how to successfully manipulate cortical excitability for the formation of new memories or the acquisition of new skills could fill an important gap between phase I and II clinical studies. TDCS studies have shown that anodal and cathodal tDCS delivered over the dorsolateral prefrontal cortex facilitate visual working memory (3). Conversely, cathodal stimulation had a detrimental effect on short-term auditory memory performance (6). Regardless of polarity, tDCS over the cerebellum disrupts practice-dependent improvement during a modified Sternberg verbal workingmemory task (8), while intermittent bifrontal tDCS impairs response selection and preparation in the same task (10). Moreover, anodal tDCS to the anterior temporal lobes delivered before the encoding and retrieval phase was effective in reducing false memories, while maintaining veridical memories (13). Finally, the application of anodal tDCS during slow-wave sleep improved declarative memory consolidation (15). Further effects of tDCS on cognitive functions in healthy individuals have been shown for decision-making (16, 17), probabilistic classification learning (21), attention (2224) and language (25, 26). Overall, these studies focused on the short-term improvements in performance induced by a single session of stimulation, typically delivered on-line during the task or immediately before it. The main limitations are the lack of control conditions over different cortical areas and the lack of a systematic monitoring of the duration of the effects. The effects of repeated applications of tDCS, their interaction with specific learning stages and tasks and the extent to which these performance improvements are retained in the long term remain to be addressed. Hypothesis-driven behavioral paradigms or stimulation strategies are also necessary to further explore the functional role of different cortical areas in human learning.

3. The clinical research of transcranial direct current stimulation

3.1 Studies in non-humans (Preclinical)

Previous animal studies have assessed safety limits of tDCS current intensity. In one study, 58 rats received tDCS with varying current densities for up to 270 minutes and histological evaluation was conducted to assess neuronal lesion. Results suggest that brain lesions occurred when current density was at least two orders of magnitude higher than typically used in humans (94) and may reflect increase in brain temperature never observed using conventional tDCS protocols (35). Another interesting insight from this study is that duration of tDCS only becomes a safety issue when the intensity of stimulation is near the threshold associated with neuronal lesion. Other animals studies conducted with different goals have also shown that tDCS used with charges similar to human studies do not induce histological lesions (95).

Finally, animal studies are useful for test dosing and exploring physiological aspects of tDCS mechanisms. On the other hand, such studies are rare, and positioning of the electrodes as well as different cortical architecture, might be critical. Still, animal models might be important for answering specific questions not possible to be done in humans.

Box 2 – Two types of study design in tDCS: “Online” versus “Offline”.

Clinical researchers usually apply tDCS in two main modalities regarding the time point in which the primary outcome variable is collected. When tDCS and the main outcome are coincident in time (i.e., when the variable is collected during tDCS application) the experiment is said to test the “online” effects of tDCS. The concept is also used when another intervention (usually having a similar time span than tDCS such as physical therapy) and tDCS are applied simultaneously. The rationale for an “online” approach is to take advantage of the putative property of tDCS to induce excitability modifications of the brain (which is analogous to TMS) to test neuromodulatory effects on the study hypothesis, such as alterations of brain functions during tDCS. For instance, an area of investigation that uses this approach is transient modulation of moral judgment and decision-making during tDCS (see discussion on ethics in this manuscript).

On the other hand, when tDCS and the variable being measured can be distinguished in time, it is said that the experiment is applying tDCS in an “offline” protocol. An “offline” tDCS protocol applies, for instance, when one surrogate outcome (or clinical parameter) is used before and after stimulation to index tDCS effects (see discussion on surrogate outcomes). An “offline” approach is also used in phase II/III tDCS studies. In such cases, tDCS is an experimental intervention and its long-term, neuroplastic effects are indexed with one or more surrogate and/or clinical outcomes.

3.2 Studies on healthy volunteers (Phase I)

In drug-based trials, phase I studies are non-randomized, non-controlled clinical (human) trials designed to address safety and optimal dosage of drugs. This is done by assessing the adverse effects/safety and dosage or the drugs. In this section, previous tDCS studies that address these questions and present issues that remain unsolved (dose parameters was above discussed) are reviewed (Table 2).

Table 2.

Main issues in the clinical research of transcranial direct current stimulation and possible solutions to effectively handle them.

Current Evidence Key issues Possible solutions
Phase I studies
Safety / Adverse Effects TDCS is not likely associated to long-term, deleterious effects. AEs are mild and transient at usual doses. Safety has not been sufficiently investigated in people with skull defects and/or patients with neuropsychiatric disorders. Further research should actively investigate adverse effects; long-term follow-up; modeling studies.
Dose- effect curve Higher doses, higher current densities and higher periods of stimulation seem to be associated with effects of larger magnitude and duration. Great between-subjects variability of effects; using higher doses is limited due to AEs; pharmacotherapy alters dose-effect curve; optimal parameters not yet defined. Further research addressing pharmacological modification of tDCS effects; increasing duration span of tDCS to avoid skin damage; bayesian approaches and modeling studies to define optimal dose.
Phase II/III studies
Subjects
Recruitment TDCS is still on its infancy, and few patients and physicians are aware of this novel technique. Non-referral due to lack of knowledge/suspiciousness of tDCS and time constraints in ambulatory settings; ethical issue of receiving placebo. Using multiple referral sources; specific neuromodulation ambulatories; building trust with volunteers and physicians (lectures, web sites, explanatory videos)
Eligibility Sample should be homogeneous, especially in phase II studies. Sources of heterogeneity are: concomitant use of medications, incorrect diagnosis of neuropsychiatric condition, wide spectrum of severity and refractoriness. Stratification during randomization; post-hoc analysis controlling for severity, refractoriness and medications; drug washout.
Attrition High attrition rates might lead to negative findings; especially if intention-to-treat analyses are performed. Daily visits to the research center and skin damage are specific issues related to dropout in tDCS trials. Careful explanation of study objectives and possible side effects; covering of transportation costs; flexible schedules; using run-in period to identify non-committers.
Methods
Blinding Blinding is the strongest approach to minimize bias. Sham TDCS involves applying an electrical current for less than 30 seconds, as to mimic intial side effects. Several studies suggested that the sham method is reliable, at least in healthy volunteers, with intermediate-high doses and in one-session studies. TDCS device can be turned off manually (single-blinded, requiring another person to evaluate subjects) or automatically (double-blinded). Further studies should explore whether this sham method is reliable in other contexts, e.g. daily stimulations for 5–10 days, higher doses and non-naïve subjects. Staff blinding should also be more carefully evaluated.
Approach To induce long-lasting (days to weeks) effects, tDCS must be delivered continuously (usually daily for 5 to 10 days) Number of sessions and time period between stimulations are still undefined as well as the extent of such effects after the initial sessions. Long follow-up of subjects (months to years); performing specific studies designed to evaluate cumulative changing in cortical excitability according to the number of stimulations (and time period between them)
Control group In tDCS research, the control group might be either a sham-group or an active group in which polarities are inverted. The latter approach is an even more reliable blinding method than sham; although it can as well induce effects. Studies exploring mechanisms of tDCS could have three groups; studies using tDCS as treatment should prefer using a sham group.

TDCS= transcranial direct current stimulation; AE = adverse effects.

3.2.1 Safety/Toxicity

Although tDCS differs in many aspects from other non-invasive neuromodulatory therapies in that it does not induce neuronal action potential and uses weak electric currents, there are safety concerns that must be addressed. If the electrochemical products generated by these currents contact the skin, skin irritation may occur; in addition, tissue heating associated with non-intact skin (therefore this is especially important in people with skin diseases and/or in protocols using daily tDCS applications and/or high electric currents) may induce skin burning (96) – although mild redness is more likely related to local, vasodilatation skin changes rather than skin damage (97). In fact, considering there is no direct contact between the brain and the electrode and also the distance, electrochemical or heating lesions to the neuronal tissue is less likely. Moreover, experimental and modeling studies suggest no significant temperature increases for typical tDCS protocols (28, 89, 91).

TDCS has been tested in thousands of subjects worldwide with no evidence of toxic effects to date. In addition to the hundreds of studies exploring tDCS effects in diverse contexts, some studies have focused specifically on safety. For instance, in a large retrospective study, Poreisz et al. (98) reviewed adverse effects in 77 healthy subjects and 25 patients who underwent a total of 567 1mA stimulation sessions. Results show the most common effects were mild tingling sensations (75%), light itching sensation (30%), moderate fatigue (35%), and headache (11.8%); and most of these effects did not differ from those of placebo stimulation. In another study, 164 sessions of stimulation were analyzed. Authors found only mild adverse effects with a low prevalence (0.11% in active and 0.08% in sham stimulation group) (99). Other initial studies (4, 5, 100103) also reported only mild, benign, and transient side effects. In fact, the most severe adverse event reported is skin lesions on the site of electrode placement (96).

Historically, the most severe adverse effect was observed in the first study of tDCS. During the 1960s Lippold and Redfearn (84) related a brief respiratory and motor paralysis in a bifrontal electrode montage with the current reference placed on the leg. No loss of consciousness was reported and respiration returned to normal when the current was stopped. This was attributed to the fact that the subject received ten times the intended intensity, probably 3mA (47).

General exclusion criteria for non-invasive brain stimulation also apply for tDCS. Subjects must be free of unstable medical conditions, or conditions that may increase the risk of stimulation such as uncontrolled epilepsy; although epileptic seizures have not been observed in a pilot study with patients with active epilepsy (104). Also, subjects must have no metallic implants near the electrodes.

Finally, it should be underscored that most of these observations were extracted from single stimulation studies in healthy subjects without medications. Less is known about the adverse effects of daily (or even twice daily) tDCS in patients with neuropsychiatric disorders who use pharmacotherapy. In such conditions, the adverse effects can be magnified and therefore they should be actively inquired during trials. For instance, some single-patient studies report that tDCS can induce mania/hypomania in patients with major depression (105107). Therefore, we suggest that a medical monitor should supervise tDCS treatment in such contexts of increased risk of significant adverse effects.

3.2.1 Dosage

TDCS dosage is defined by the following parameters: (1) current dosage (measured in amperes); (2) duration of stimulation; and (3) electrode montage (size and position of all electrodes). Current density (current dose divided by electrode size) is also an important parameter in considering dosage; especially for defining safety - (77); see the following review for additional information (108). The most common electrode sizes are of 25–35 cm2 with currents of 1–2 mA (generating densities ranging from 0.28–0.80 A/m2) for up to 20–40 minutes. However, the current that effectively reaches neuronal tissue depends on other less controllable factors. These include skin resistance, skull resistance, resistance of intracranial structures (e.g. blood vessels, cerebrospinal fluid, and meninges) and the resistance of brain tissue, which varies according to cell type and structure (e.g. glial cells, pyramidal neurons, white matter, and so on). Moreover, patients with skull defects, brain lesions and other conditions will influence current amount and delivery. In addition, the baseline cortical excitability is different in people using pharmacotherapy (e.g. benzodiazepines (74), anticonvulsants (109 2006), antidepressants (79) and others (110)) and/or presenting neuropsychiatric disorders (e.g. major depression (111), schizophrenia (112), fibromyalgia (113), migraine (114), and others); an issue that is likely to interfere with the chosen dosage. Finally, other variables influence baseline cortical excitability such as gender (115), age (116) and smoking (117). Hence, the same amount of current is likely to have non-uniform effects in subjects with different conditions. For instance, one study showed that low (25mg) and high (200mg) doses of l-dopa abolished tDCS-induced effects on cortical excitability whereas an intermediate (100mg) dosage increased inhibitory effects (118). Notwithstanding, these studies should be regarded as exploratory and thus replicated in other contexts and samples; especially in clinical populations.

Therefore, in the context of clinical research, such individual factors are a source of variability and, if important enough, may result in negative findings. In order to avoid this, one alternative is to standardize the source of error in the sample. For instance, using saline-soaked sponges to minimize skin resistance (which can also be measured by an ohmmeter adapted in the tDCS device – some devices do give the resistance), excluding patients under pharmacotherapy, or controlling when it is not feasible (e.g. benzodiazepines), avoiding sample heterogeneity using specific diagnostic criteria, particularly when working with a small, neuropsychiatric subject pool. Future studies addressing the interaction of tDCS and drugs in psychopharmacology will continue to explore and identify which drugs do not interfere with tDCS effects and which ones could block or enhance tDCS-excitability effects (31, 47).

Initial studies measuring brain excitability demonstrate that currents as low as 0.28 A/m2 present depolarizing and hyperpolarizing effects (27, 28). Additional phase I/II studies addressed the effects of varying dose and/or time of stimulation on cortical excitability and/or neuropsychological tasks. Ohn et al. (119) tested the effects on working memory during 30 minutes of stimulation, showing that performance increased in a time-dependent fashion. Other studies showed the cognitive effects induced by tDCS are dependent on the current intensity; demonstrating effects such as enhanced verbal fluency improvement at 2 mA (vs. lower improvement at 1 mA) (38); and working memory improvement at 2 mA (vs. no improvement at 1 mA) (120). Nevertheless, it remains unclear whether there is a linear (dose versus effect) curve associated with direct current stimulation and the influence of each parameter (dose, current density, stimulation duration) on these effects. It is known that increasing current densities will increase the depth of the electrical field, thus affecting different populations of neurons. However, at greater intensity tDCS might be painful to the subjects. For these reason, a more effective approach designed to prolong tDCS effects is to increase the stimulation duration as opposed to the current density (28, 47, 55, 57).

Short applications (i.e., seconds to few minutes) of anodal/cathodal tDCS result in excitability shifts during stimulation but no after-effects. However, no long-term effects are seen. In contrast, ten minutes or more of stimulation can elicit prolonged after-effects, which can be sustained for over an hour (28, 47, 59). The exact duration of effects depends on the targeted cortical area and on the type of variable assessed.

For clinical purposes, longer-lasting effects are crucial. Single-dose tDCS interventions have relatively short-lived after-effects. Multiple stimulation sessions are required to induce a significant manipulation in synaptic efficacy (121, 122). In fact, repeated sessions of tDCS may have cumulative effects associated with greater magnitude and duration of behavioural effects. For example, cathodal tDCS applied over 5 consecutive days is associated with cumulative motor function improvement lasting up to 2 weeks after the end of stimulation. This is an effect which is not observed when sessions are applied weekly (as opposed to daily) (102). Whether this approach is appropriate to maximize and stabilize the electro-physiologic effects of tDCS remains under investigation. The optimal repetition rate and duration to promote tDCS-induced plasticity also remains to be determined. In animal experiments, repetition of tDCS during the after-effects of a first stimulation session has been shown to enhance efficacy (52). However, repeated plasticity induction may result in homeostatically driven antagonistic effects (123). Recently, Monte-Silva and coworkers (122) directly compared the effects induced by single sessions of cathodal tDCS over the motor cortex to the effects of repetitive stimulation during or after the after-effects of the first stimulation. The results showed that increasing cathodal tDCS duration (1 mA, with no inter-stimulation interval) resulted in longer lasting after-effects, typically over 1 hour (tDCS duration from 9 to 18 min prolonged the after-effects from 60 to 90 min). Interestingly, when the second stimulation was performed during the after-effects of the first, a prolongation and enhancement of tDCS-induced effects for up to 120 min after stimulation was observed. In contrast, when the second session was performed 3 or 24 hours after the first, tDCS effects on cortical excitability were mixed. This was shown with a primary reduction or abolishment of the initial effects of cathodal tDCS, followed by a later re-occurrance of tDCS-induced cortical inhibition. Such neurophysiological evidence is indicative of a stimulation timing–dependent plasticity regulation in the human motor cortex. Understanding the interaction of the consecutive stimulation protocols appears crucial to effectively target spontaneous changes of cortical activity and excitability. Hence implementing more effective procedures of plasticity induction procedures in clinical settings is crucial – in fact these results need to be replicated in clinical populations

3.3 Studies on patients with neuropsychiatric conditions (Phase II/III)

Phases II and III studies relate to using an intervention in clinical samples. Phase II studies are typically small and use targeted samples to obtain additional information regarding optimal parameters of stimulation. Phase III are pivotal studies, involving larger samples. In the US, two positive phase III trials are required for approving a drug or device by the Food and Drug Agency (FDA).

As mentioned above, several studies have explored the therapeutic application of tDCS in several neuropsychiatric disorders. The results of these studies reveal long-lasting tDCS effects and have promoted its use in clinical settings. Because clinical development of tDCS is being conducted mainly in Academia, studies are not widely standardized regarding variables and population samples; therefore limiting conclusions. These findings are also limited by small sample sizes and experimental design. In fact, a similar scenario has been observed for TMS five to ten years ago (124). This section aims to comprehensively review the main issues of the later phases of clinical trial development for tDCS.

3.3.1 Issues related to recruitment and eligibility

Recruiting subjects for tDCS clinical trials presents a challenge. Proposing an intervention alternative to the mainstream pharmacotherapy might be seen by prospective patients and referring physicians in non-academic settings as suspicious. This issue can be particularly important when a large sample size is required and/or if the eligibility criteria exclude refractory patients who are more prone to enroll in research protocols. Likewise, referral physicians, due to time constraints in the ambulatory setting, usually prefer to treat drug-naïve patients themselves. Indeed, daily visits for one-to-two weeks to research centers might sound unappealing and/or unaffordable even to refractory patients. In such contexts, it is advisable to have multiple referral sources and to use broad recruitment strategies. Building trust with potential volunteers is imperative. One cost-effectiveness approach could be using explanatory videos in lay language (46).

Another issue is sample heterogeneity. In pivotal clinical trials comparing tDCS against pharmacotherapy, large samples are typically required and patient heterogeneity might be larger than for drug trials for the same condition. This is due to the fact that the severity of the condition ranges from drug-naïve to refractory subjects. Targeting only the former would create difficult enrollment (for the reasons mentioned above), while targeting only the latter decreases overall generalizability. Possible solutions include stratification during randomization (refractory vs. non-refractory), post-hoc analysis controlling for refractoriness, or increasing sample size to address some of the issues associated with heterogeneity.

3.3.2 Blinding issues

It appears easier to conduct sham-controlled trials using tDCS compared to TMS. TMS induces itching and pain sensations over the stimulation site whereas tDCS induces a mild tingling sensation that usually rapidly fades. Therefore, sham protocols begin with active tDCS, which is switched off within a minute. In addition, the tingling sensation relates to the velocity in which the current is either increased or decreased. In fact, an increase of current delivery from 0.1 to 0.2mA/sec generates no discomfort for most subjects (125). Interestingly, some subjects in the sham group continue feeling some tingling even after the current is discontinued.

These sensations are related to the total amount of charge delivered. Although this has yet to be systematically evaluated, this relationship can be a potential issue when delivering relatively high charges (>1.5–2mA/sec) and/or higher current densities. There is evidence that electrolyte solutions with lower NaCl concentrations (15 mM) are perceived as more comfortable during tDCS than those solutions with higher NaCl concentrations (220 mM). Since the ionic strength of deionised water is much less than that of all NaCl solutions, there is a significantly larger voltage required to carry current through the skin compared to NaCl solutions. Thus, it is recommend the use of solutions with relatively low NaCl concentration, in the range 15 mM to 140 mM, as tDCS at these concentrations is more likely to be perceived as comfortable, requires low voltage and still allows good conduction of current.(126). It has also been proposed to apply topical anesthetics to alleviate this issue (47).

An additional blinding issue is the local vasodilatation after tDCS. This causes the skin to turn red that might not be acknowledged by the subject but might be seen by the staff and other patients. In clinical protocols, such redness can be evident after several days of stimulation. This can become a logistical issue, demanding stimulated patients to leave the setting immediately, avoiding contact with other people (patients and researchers) as to avoid blinding breaking. Another approach would be to interview patients before (and not after) being stimulated. If a rater notices evidence of redness on the scalp of a patient, another blinded rater should substitute him/her, although this matter is more important in sham-controlled studies as in studies using active groups differing only regarding polarity (and not scalp site of stimulation) cathodal and anodal stimulation cannot be distinguished between each other. Also 30 seconds of active stimulation in the sham protocols might also lead to local redness.

Clinical protocols should assess post-hoc the effectiveness of blinding; though investigators need to be aware that potential differences might occur because active tDCS is more effective than sham tDCS. It is not easy to detangle unblinding vs. response because effectiveness. Other alternatives are (1) to avoid crossover trials, especially when the crossover happens in the same section, as to avoid subjects noticing the differences; (2) to apply active protocols but switching polarity so that adverse effects do not threaten blinding even if noticed –although here the issue would be whether changing polarity would be an appropriate control condition, when the reference electrode is not physiologically inert.

3.3.3 Study design

Four approaches in tDCS clinical trials for neuropsychiatric disorders are possible: (1) to compare active vs. tDCS sham in a superiority trial; (2) to compare tDCS vs. another therapy (e.g. acupuncture, pharmacotherapy) as a superiority or non-inferiority trial; (3) to combine tDCS with another therapy (e.g. physical therapy, pharmacotherapy) vs. sham tDCS and another therapy as a superiority trial, and (4) combination of these approaches.

Two-arm designs are suitable when comparing active vs. tDCS sham, an approach commonly used in pilot, “proof-of-concept” studies. This approach is effective in studies exploring the mechanisms of action of tDCS, e.g., with neuroimaging or serum measurements.

Three-arm and “double-dummy” (i.e., placebo pill + active tDCS vs. pharmacotherapy + sham tDCS) designs are adequate for comparing tDCS against another therapy. The placebo arm is interesting for increasing assay sensitivity, although ethical concerns might impede using placebo groups when there are reasons to believe that treatment efficacy among study arms is imbalanced (principle of clinical equipoise) (127).

Another option is a factorial (2 × 2) design, which could be useful to test tDCS with and/or against another therapy of interest. For instance, in a trial testing tDCS for chronic pain, patients could be randomized to four groups: only tDCS, only pharmacotherapy, tDCS and pharmacotherapy and sham plus placebo. In fact, such a design is the most robust as it tests two interventions simultaneously and also one intervention against another, making them optimal for pivotal studies. Although comprehensive, this approach is more demanding regarding resources, sample size, and logistics.

The n-of-one (n=1) trial is a possible approach when the researcher is confident that tDCS effects are short-lasting (which is not usually the case for studies employing multiple sessions of tDCS). In this design, one subject is randomized to receive repeated randomized allocations of the tDCS treatment. This is helpful especially to address different parameters of stimulation for single session protocols.

3.3.4 Attrition

Attrition (or “dropout”) is the premature discontinuation of participation in a trial occurring either immediately after the baseline visit or at any time before endpoint. The specific reasons for attrition in tDCS trials should still be investigated. Although some might be the same for pharmacotherapy, one reason more specific for tDCS trials is the difficulty to comply with required daily visits to the research center (that usually occur during the first two weeks of the study). In intention-to-treat trials, this issue can be particularly perturbing as such subjects will maintain the same baseline scores at endpoint and thus diminish the effect size between groups. To avoid attrition in tDCS trials, some measures can be taken such as: (1) concede one or two non-consecutive missing visits, which are replaced at the end of the daily stimulation phase and; (2) using a “run-in” period, that is, a phase before trial onset in which subjects receive either active or placebo/sham treatment (usually for one week) as a method to preemptively screen and discard non-adherent subjects. Although the usefulness of run-in phases is controversial in pharmacotherapy given the potential for selection bias, the rationale for using tDCS is to select subjects that can commit to the stimulation protocol requirements.

Finally, although uncommon, another issue is skin burn. This would prevent further stimulations, breaking blinding and also forcing the investigators to withdraw treatment, leading to a study dropout. Skin burning can be avoided by diminishing electric density (i.e., increasing electrode size and/or diminish electric current) and electric resistance (by using rubber electrodes involved with saline-soaked sponges) over the stimulation site.

3.3.5 Statistical issues

Being that most tDCS trials are exploratory and employing small samples, they are particularly vulnerable to type I and type II errors.

Type I (false-positive) errors occur in exploratory studies performing several statistical tests, being the case of many phase-II tDCS trials. In this scenario, investigators need to decide whether to claim findings as exploratory or to determine a priori the statistical method for the primary outcome, differentiating other statistical analyses as secondary.

Type II (false-negative) errors occur in small studies and are related to underpowered trials. Again, most phase-II tDCS trials recruit small samples and are prone to this error. To avoid this, researchers must perform sample size calculations when designing the trial. Another approach is to employ adaptive designs, which allow sample increasing during the study. Although, this method may be challenging for researchers and readers to interpret the data. In this context, given that most of tDCS trials are conducted with limited resources, the best choice of primary study outcome is a continuous outcome and two time points so as to increase statistical power (and consider other analyses as secondary). Although baseline differences are usually not significant in tDCS trials probably because trials have a relatively homogeneous population, one option here is to calculate normalized differences from baseline. In this case a simple approach to calculate sample size is to use independent two-sample t-test provided in most statistical software packages.

3.3.5 Pilot versus Pivotal studies for transcranial direct current stimulation

Most phase II studies are also referred as “proof-of-concept” or “pilot” studies. These studies typically use small, high-targeted samples that represent the more severe spectrum of a disease to address the efficacy of a given treatment in optimal conditions. They also use several surrogate endpoints and perform many exploratory analyses. Exploratory phase II studies are necessary as they provide data to be used in subsequent trials. Furthermore, data of small studies can be pooled together in meta-analyses. However, the validity of these analyses can be contested when approving clinical interventions (128, 129).

An additional challenge for pilot studies is the exploratory nature and thus an important degree of risk regarding outcomes that is normally not seen in animal models in neuromodulation research. This hinders the ability to test the clinical efficacy of tDCS for a particular condition for the first time. In such context, a negative finding might be due to tDCS parameters or a poor neurobiological model (e.g., a negative finding in a pain trial with anodal stimulation over the dorsolateral prefrontal cortex area might represent – besides being a true-negative - either the use of incorrect tDCS parameters, a misconception in the neurological model; thus the dorsolateral prefrontal cortex area being unrelated to pain pathophysiology). This issue poses an additional challenge in tDCS research.

Therefore, pivotal (phase III) studies are necessary to validate tDCS as an effective treatment when proof-of-concept trials showed encouraging results. Future phase III studies should include: (1) sample size estimation based on prior, pilot trials or meta-analyses; (2) robust blinding method (example: using tDCS devices that can be automatically turned off as to keep both patients and appliers unaware of the intervention delivered) and (3) assessment of sample heterogeneity, either targeting particular samples (e.g. medication-free patients) or identifying potential sources of heterogeneity (e.g., degree of refractoriness, number of depressive episodes, depression severity, and others) and controlling for them during study design (stratified randomization approaches) or statistical analysis.

Box 3 - Insights from tDCS studies for major depression.

In the past ten years, several trials applied tDCS to subjects with major depressive disorder (MDD). Fregni et al. (1) performed a pilot randomized, sham-controlled, double-blind trial in which 10 patients were randomly assigned to receive either 5 days of active or sham stimulation. Boggio et al. (2) also enrolled 40 MDD subjects with different degrees of refractoriness (but medication-free) and randomized them to 10 sessions of active DLPFC tDCS, active occipital tDCS or sham tDCS. The findings suggested that the active DLPFC tDCS group presented a superior, significant improvement in HDRS scores compared to the other groups. Rigonatti et al.(4) demonstrated in an open-label study that Fluoxetine 20mg/day and active tDCS (from patients of Boggio’s study) presented similar scores after 6 weeks of treatment. Ferrucci et al. (5) stimulated 14 patients with severe MDD using 2mA for 20 minutes for 5 days twice a day, showing a significant improvement in mood. Such effects seem to be more robust in more severe patients (7 2009). Loo et al. {enrolled 40 patients with severe MDD, in a double-blinded, sham-controlled study but failed to demonstrate significant difference between groups in this phase; tDCS was only more effective during the open-label phase in which patients received additional 5 sessions. However, this study has some limitations: the dose applied was relatively low (1mA), and only 5 stimulations sessions were held, which were alternated (other studies used consecutive sessions). Moreover, patients with axis II disorders were not excluded. Finally, Brunoni et al. (12) compared patients with unipolar vs. bipolar depression and found that tDCS might be a potential treatment for both conditions. However, as with phase II trials, these studies share common characteristics: relatively small sample sizes, heterogeneous sample (e.g., refractoriness, medication use), blinding vulnerability (some studies were open-label), absence of primary hypothesis (most of them used several depression rating scales), and presence of “carryover” effects (in crossover studies). These initial trials likely incurred in some false-positive and false-negative results; nevertheless, they reveled the potential effectiveness of tDCS for major depressive disorder. Finally, a search made on clinicaltrials.gov in September 2010 revealed that there are at least seven trials exploring the antidepressive effects of tDCS worldwide; and the design and methods of one of them (18) has been recently published.

3.3.6 Surrogate outcomes

Although several definitions for surrogate (or substitutive) outcomes exist, they are typically understood as laboratory measurements that substitute clinically meaningful outcomes for being in a prior step in the pathophysiological pathway of the disease (130). In neuromodulation research, this also includes neuropsychological tests and neuroimaging scans. The advantage of using surrogate outcomes is avoiding long-term, expensive research. This is achieved by substituting “hard” outcomes (death or serious events) for “soft” measurements that take place earlier. Furthermore, surrogate outcomes must have high accuracy and low variability; otherwise their utility is limited (Table 3).

Table 3.

Substitutive outcomes used in transcranial direct current stimulation research.

Outcome Definition Pros Cons
Neuropsychological tests Paper or computerized tests that explore cognitive performance. Non-expensive, relatively easy to apply, specific tests can be used according to the brain area under study. Low signal-to-noise ratio, as performance depends on the rater, the subject's characteristics (age, educational level); learning effects, thus requiring a control group. Relatively non-specific.
Neurophysiological measurements Techniques that record electrical brain activity (EEG, qEEG,ERP) as to examine changes in brain activity. Strong temporal relationship (i.e., very sensible to change), able to index subclinical changes. Relatively non-specific (measurement of several brain networks) and medium to low spatial relationship. Devices should be adapted to minimize electrical noise of tDCS.
Transcranial Magnetic Stimulation (TMS) TMS used as a tool to index cortical excitability Relatively affordable and easy to apply, provide several measures of cortical excitability. Measures obtained with TMS have considerable intra and inter-subject variability; usually applied over the motor cortex only.
Neuroimaging methods - Radiotracers Methods such as PET and SPECT that use radioactivity to assess brain metabolism. Good temporal relationship, able to index subclinical changes, can be used during "online" tDCS. Poor spatial resolution, invasive, PET is expensive and not always available (requires a cyclotron).
MRI-based neuroimaging methods MRI-based analyses (e.g. functional, structural, spectroscopy, DTI) that provide static and dynamic neuroimages. Not-invasive, excellent spatial and temporal resolution (according to the method), specific. Analyses are difficult and can yield false-positive results, expensive, not always available, tDCS devices and MRI cannot be used simultaneously.
Blood chemistry Blood measurement of substances expressed by the CNS that cross BBB. Minimally invasive, easy to perform, samples can be frozen and analyzed later, gives a quantitative measurement, sensible to change. Minimal spatial resolution (brain metabolites are usually not specific of a particular area), low temporal resolution (metabolites must cross the BBB), lack of important biological blood markers in neuropsychiatry.

tDCS= transcranial direct current stimulation; EEG= electroencephalography; qEEG= quantitative EEG; ERP = evoked-related potentials; PET = positron-emission tomography; SPECT = single photon emission computed tomography; MRI = magnetic resonance imaging; DTI = diffusion tensor imaging; CNS = central nervous system; BBB= blood-brain barrier.

One surrogate outcome that is often used is TMS-indexed cortical excitability, a neurophysiological measurement. According to the protocol utilized, it indexes and detects changes in brain activity (131). For instance, measurement of motor threshold – the lowest intensity to elicit motor-evoked potentials of more than 50 uV in at least 50% of trials – is used for studying whether different tDCS protocols change motor cortical excitability. Also, measurement of the silent period – the period of electromyographic suppression (or voluntary muscle activity) after one single suprathreshold TMS pulse – can be also used for addressing whether and how tDCS affects the inhibitory cortical interneurons that are recruited during this task. Moreover, paired-pulse TMS is also used for studying inhibitory or excitatory cortical mechanisms elicited after one suprathreshold pulse and is another method that can be coupled with tDCS for indexing cortical excitability. Nonetheless, all these methods are limited to the motor cortex and thus might not necessarily reflect net brain cortical excitability and/or cortical excitability of specific brain areas.

Neuropsychological tests are able to measure brain activity in some areas, especially those that cannot be indexed through TMS. Moreover, cognitive deficits are a common consequence of brain injury, stroke, epilepsy, neurodegenerative and other neurological disorders. Hence the rehabilitation of cognitive function, such as language, spatial perception, attention, memory, calculation, and praxis represents an expanding area of neurological rehabilitation and has recently attracted growing attention within the scientific community. For instance, changes in the activity of the prefrontal cortex can be measured using tests of working memory and attention, while temporoparietal stimulation can be evaluated using working memory tests. A drawback of several neuropsychological tests is the need of a control group to adjust for learning effects biases. Performance is also influenced by educational level and, therefore, the results of one study might not be valid for similar samples in different countries.

Neurophysiological measurements are another possible approach to surrogate outcomes. Besides TMS, brain activity can be measured using electrodes, which can be interpreted using several methods. These include the qualitative EEG, which measures spontaneous neuronal firing; the Event-Related Potentials (ERPs), which modifies according to the brain area provoked; the quantitative EEG (qEEG) which maps brain activity; and, finally, new approaches that provide a three dimensional brain imaging based on electromagnetic reconstruction of the brain (which in fact are not widely accepted due to the “inverse problem solution”. For a review on this topic, see Pascual-Marqui et al. (132)). Such measurements lack specificity –simple psychological, cognitive, or motor task recruits several brain networks and thus the measured ERP can be an epiphenomenon of another brain region rather than a relevant finding (i.e., a “noise” and not a “signal”). Another issue is that the devices measuring brain activity must be adapted in order to decrease the electrical noise generated by the tDCS device; or, alternatively, the measurement must be collected either before or after (but not throughout) tDCS delivery.

Neuroimaging methods are divided into two branches: the first uses radiotracers and is represented by the positron-emission tomography (PET) and the single-photon emission computed tomography (SPECT), which assess brain metabolism through the emission of gamma ray. The advantage of PET/SPECT in tDCS research is that the radiotracer can be injected during brain stimulation, thus providing “real-time” brain imaging. However, the spatial resolution of such methods is poor. Because they obligatory require using radiotracers, the radiation dose needs to be carefully controlled and monitored. The second branch of neuroimaging is the magnetic resonance imaging (MRI). This technique presents high spatial resolution. There are several methodological approaches for MRI, which allows evaluation of different aspects of brain activity. For example, functional MRI (fMRI) explores the paramagnetic properties of hemoglobin to infer brain metabolism (based on blood oxygen saturation) while magnetic resonance spectroscopy (MRS) analyzes the magnetic fields of relevant molecules (e.g. glutamate, GABA) and provides a noninvasive “chemical biopsy” of the brain. Some of these techniques such as fMRI lack temporal resolution as it does not measure electrical activity changes directly (it does indirectly via changes in cerebral flow). Diffusion tensor imaging (DTI) focuses on the white matter fibers, revealing the neural connectivity between brain areas. Finally, voxel-based morphometry (VBM) is a computational analysis of morphological images that makes inferences about brain activity based on the differences of brain tissue concentration among areas. For tDCS, these methods present the advantage of high spatial resolution; allowing to assess subtle changes in the stimulated area. For instance, one study used VBM to assess neuroplastic changes after five days of TMS over the superior temporal cortex; showing macroscopic gray matter changes in the region (133). Even though, the reliability of some methods of MRI are currently under dispute (134). Moreover, tDCS is not used concomitantly with MRI yet due to serious risks of overheating and thus an “online” visualization of the stimulated area is not possible although this technical difficulty might be resolved in the near future.

Finally, there is a wide range of blood measurements used in neuropsychiatry research for surrogate outcomes (135). One biomarker under intensive investigation is the brain-derived neurotrophic factor (BDNF). This marker plays an important role in synaptogenesis and neuroplasticity and is thus believed to be linked with some neuropsychiatric disorders – for instance, BDNF serum levels are low in depressed patients and increase after antidepressant treatment (136). A recent study showed BDNF expression also increases after tDCS (29). Additional biomarkers used in neuropsychiatry include inflammatory proteins such as interleucin-1, interleucin-6 and TNF-alpha (137); hypothalamic-pituitary-adrenal activity, which is measured by serum and salivary cortisol (138); and oxidative stress proteins such as nitric oxide and other neuroinflammatory protein markers (139, 140). These biomarkers present two important drawbacks: first, due to the blood-brain barrier, serum levels might not reflect “real-time” brain activity (or even brain activity at all); second, serum levels can only express the net brain activity, and do not represent a specific area. Therefore, perhaps the most effective use in tDCS research is to index disease improvement in phase II/III studies.

Box 4 – Challenges for outcome measures in tDCS clinical research.

As neither the full spectrum of clinical efficacy nor the mechanism of action of tDCS are completely described, outcome measures for tDCS trials ideally will inform both about tDCS clinical potency and about the biology of tDCS. With respect to clinical data, the common accepted behavioral outcomes might be insensitive to subtle changes in neurologic function. This is particularly relevant for tDCS as it has a modest (perhaps subclinical) neuromodulatory and behavioral effect, particularly for single exposures. Thus in the present early stages of investigation, the field of study may benefit from clinical trial designs that incorporate secondary outcomes in addition to measures of the patient’s chief symptom. Among these are changes in normal function that may be affected by tDCS. For example, an investigator testing tDCS effects on chronic pain might add a battery of motor tasks to see whether there is any subtle loss of normal function with treatment. Similarly, an investigator applying tDCS for treatment of epilepsy may add a questionnaire to assess mood.

Further, prospects for improving tDCS clinical efficacy improve if the tDCS mechanism of action is better understood. To date, the common feature in tDCS trials appears to be its capacity to produce a lasting change in regional cortical excitability. Given these data, outcome measures aimed to capture the extent to which tDCS induces synaptic plasticity may also be useful additions to ongoing trials. That is, one could ask whether tDCS improved the symptom in question, and in parallel ask whether an LTP-type or LTD-type change in regional cortical excitability has occurred. If so, then perhaps in future trials, the tDCS effect may be augmented by the addition of appropriate pharmacologic agents or behavioral tasks that facilitate synaptic plasticity. As an example, in future trials in which cathodal tDCS may be applied over an epileptic seizure focus, whether LTD-type suppression has occurred over the stimulated area can be determined within hours of tDCS. However to find out whether seizures are reduced in frequency may take days to weeks. Thus subjects can be stratified into groups that have or have not undergone regional LTD, and clinical outcomes can be evaluated separately for subjects that did and did not experience regional depotentiation. This subclassification of subjects in an epilepsy trial would potentially reduce confounding results from subjects where tDCS was not biologically effective at the time it was administered. Additionally, investigators would be wise to bear in mind the potential pitfall of choosing outcome scales that are not sufficiently sensitive to capture a relatively modest clinical tDCS effect. Thus, if tDCS strongly changes a component of a larger clinical scale, further research can be stimulated, even if negative results were found initially.

3.4 TDCS in children

As the brain is under intensive development during childhood and adolescence – particularly the prefrontal cortex (141), intensive research is currently being made to explore how cognition, emotion, behavior and other functions evolve. Having neuromodulatory properties, tDCS would be an interesting tool to explore which brain areas are particularly important in each stage of development both in healthy and pathological conditions, such as epilepsy, cerebral palsy, autism and mental deficiency. However because of its potential to induce neuroplastic changes, tDCS should be used carefully especially during important phases of brain development associated with intensive plasticity and also other processes such as synaptic pruning.

A further step would be using tDCS for treating neuropsychiatric disorders in children, but this has not been tested yet. In a review of TMS studies in children, no adverse effects were reported, but its use is still limited for some reasons, including lack of established safety guidelines (142). Notwithstanding, tDCS is a promising tool for children neurology and psychiatry.

4. The ethics of transcranial direct current stimulation

TDCS is a putative candidate for adjuvant therapy for a range of neuropsychiatric conditions. tDCS is a valuable tool in neuroscience research, as its focality can be used to explore several brain aspects. Studies regarding tDCS ethics reveals its ability to induce changes in behavior such as in moral judgment (143), deception (144, 145) and decision-making (146). For instance, one recent study showed tDCS affected utilitarian behavior. Similarly to other studies in tDCS, the polarity-dependent effects resulted in a selfish vs. selfless behavior in women (143). Although the effects were short-lasting (volunteers were not exposed to daily stimulation), the targeted area is similar than used in studies exploring the long-lasting tDCS effects. Therefore, the ethical concern is whether tDCS could induce maladaptive behavior changes, and if so, to what intensity and extent of time.

Diverse tDCS studies on healthy subjects have shown positive changes in attention and memory (23, 24, 147). From the scope of neuroethics, the issue is whether tDCS enhances cognition in healthy subjects. Can tDCS be used to boost performance in specific situations (e.g. before school tests)? Another issue is thatthe cognitive effects described (increased attention and memory) from tDCS are in some aspects similar to amphetamines.Despite therapeutic applications, amphetamines are sold illegally as a recreational and performance enhancer drug (with the suggestive name of “speed”). As a tDCS device is easily built and inexpensive (contrary to TMS), it could also be used for non-research and non-therapeutic objectives by lay people. In fact, there are online videos in popular web sites such as Youtube explaining how to build and use a tDCS device (148). Although it should be underscored that all the enhancement effects were present for a short period of time, it is possible that prolonged daily stimulation could increase the time span of such effects, thus inducing maladaptive changes. On the other hand other legal substances such as caffeine are also frequently used as cognitive boosters.

In fact, because applications in these fields are currently in the research stage, fixed protocols and safety guidelines are yet to be defined. Research and development of any new devices provides an opportunity for brain science and clinical care to advance, and also challenges the medical and wider communities to address potential dangers and complications, ethical and moral quandaries, and issues of healthcare economics and distributive justice. For innovative neurotechnologies, these are major potential pitfalls to look out for. Intervening in the brain is always fraught with the potential for serious consequences. Despite these concerns, only by conducting carefully planned clinical and experimental studies can we provide the impetus to advance care for people with brain, emotional or psychological, or neuropsychiatric disorders.

5. Conclusion

The present paper addresses the main aspects of the clinical research of tDCS. This technique has a wide range of potential applications and can be used to explore the basic aspects of neurosciences as well as for the treatment of neuropsychiatric disorders. TDCS has unique characteristics such as ability to induce antagonistic effects in cortical excitability according to the parameters of stimulation; concomitant (“online”) use with neuropsychological and psychophysiological tests; non-invasiveness and thus absence of pharmacokinetics interactions, being a putative substitutive/augmentative agent in neuropsychiatry; and low-cost and portability, making it suitable for increasing access to novel therapies. However, such characteristics also bring challenges regarding clinical design, neuroethics and legal issues. In this paper, we aimed to provide an overview of transcranial direct current stimulation in clinical research; thereby providing knowledge for conducting proper clinical trials using this promising approach.

Acknowledgments

We are thankful to Erin Connors for copyediting this manuscript. We are also grateful to Scala Institute and Mackenzie University (Sao Paulo, Brazil) for the additional support to organize this working group meeting in the II International Symposium in Neuromodulation. This working group meeting was the 2nd international tDCS club workshop, which took place in Sao Paulo, Brazil, in March 2010, at the Social and Cognitive Neuroscience Laboratory of Mackenzie Presbiterian University. The first meeting was held in Milan, Italy in 2008.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Conflict of Interest: AP and RF have some stock options in the neurostimulation company Newronika srl (Milan, Italy).

Contributor Information

Andre Russowsky Brunoni, Department of Neurosciences and Behavior, Institute of Psychology, University of São Paulo, São Paulo, Brazil.

Michael A. Nitsche, Department of Clinical Neurophysiology, Georg-August University, Goettingen, Germany

Nadia Bolognini, Department of Psychology, University of Milano-Bicocca, Milan, Italy & Neuropsychological Laboratory, IRCCS Instituto Auxologico Italiano, Milan, Italy.

Marom Bikson, The City College of City University of New York, New York, USA.

Tim Wagner, Massachusetts Institute of Technology, Boston, USA.

Lotfi Merabet, Massachusets Eye and Ear Infirmary, Harvard University, Boston, USA.

Dylan J. Edwards, Burke Institute of Medical Research, White Plains, NY, USA

Antoni Valero-Cabre, Boston University School of Medicine, Boston, MA, USA.

Alexander Rotenberg, Department of Neurology, Children’s Hospital, Harvard Medical School, Boston, MA, USA.

Alvaro Pascual-Leone, Berenson-Allen Center for Non-invasive Brain Stimulation, Beth Israel Deaconess Medical Center, Harvard Medical School, Boston, MA, USA.

Roberta Ferrucci, Centro Clinico per la Neurostimolazione, le Neurotecnologie ed i Disordini del Movimento, Fondazione IRCCS Ca’ Granda Ospedale Maggiore Policlinico, Università degli Studi di Milano Dipartimento di Scienze Neurologiche, Milan, Italy.

Alberto Priori, Centro Clinico per la Neurostimolazione, le Neurotecnologie ed i Disordini del Movimento, Fondazione IRCCS Ca’ Granda Ospedale Maggiore Policlinico, Università degli Studi di Milano Dipartimento di Scienze Neurologiche, Milan, Italy.

Paulo Boggio, Social and Cognitive Neuroscience Laboratory and Developmental Disorders Program, Center for Health and Biological Sciences, Mackenzie Prebyterian University, Sao Paulo, Brazil.

Felipe Fregni, Laboratory of Neuromodulation, Department of Physical Medicine & Rehabilitation, Spaulding Rehabilitation Hospital and Massachusetts General Hospital, Harvard Medical School, Boston, MA.

References

  • 1.Fregni F, Boggio PS, Nitsche MA, Marcolin MA, Rigonatti SP, Pascual-Leone A. Treatment of major depression with transcranial direct current stimulation. Bipolar disorders. 2006 Apr;8(2):203–204. doi: 10.1111/j.1399-5618.2006.00291.x. [DOI] [PubMed] [Google Scholar]
  • 2.Boggio PS, Rigonatti SP, Ribeiro RB, Myczkowski ML, Nitsche MA, Pascual-Leone A, et al. A randomized, double-blind clinical trial on the efficacy of cortical direct current stimulation for the treatment of major depression. Int J Neuropsychopharmacol. 2008 Mar;11(2):249–254. doi: 10.1017/S1461145707007833. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Fregni F, Boggio PS, Nitsche M, Bermpohl F, Antal A, Feredoes E, et al. Anodal transcranial direct current stimulation of prefrontal cortex enhances working memory. Exp Brain Res. 2005 Sep;166(1):23–30. doi: 10.1007/s00221-005-2334-6. [DOI] [PubMed] [Google Scholar]
  • 4.Rigonatti SP, Boggio PS, Myczkowski ML, Otta E, Fiquer JT, Ribeiro RB, et al. Transcranial direct stimulation and fluoxetine for the treatment of depression. Eur Psychiatry. 2008 Jan;23(1):74–76. doi: 10.1016/j.eurpsy.2007.09.006. [DOI] [PubMed] [Google Scholar]
  • 5.Ferrucci R, Bortolomasi M, Vergari M, Tadini L, Salvoro B, Giacopuzzi M, et al. Transcranial direct current stimulation in severe, drug-resistant major depression. J Affect Disord. 2009 Nov;118(1–3):215–219. doi: 10.1016/j.jad.2009.02.015. [DOI] [PubMed] [Google Scholar]
  • 6.Elmer S, Burkard M, Renz B, Meyer M, Jancke L. Direct current induced short-term modulation of the left dorsolateral prefrontal cortex while learning auditory presented nouns. Behav Brain Funct. 2009 Jul 15;5(1):29. doi: 10.1186/1744-9081-5-29. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Ferrucci R, Bortolomasi M, Brunoni AR, Vergares M, Tadini L, Giacopuzzi M, et al. Comparative benefits of Transcranial Direct Current Stimulation (tDCS) treatment in patients with mild/moderate vs. severe depression. Clinical Neuropsychiatry. 2009;6:246–251. [Google Scholar]
  • 8.Ferrucci R, Marceglia S, Vergari M, Cogiamanian F, Mrakic-Sposta S, Mameli F, et al. Cerebellar transcranial direct current stimulation impairs the practice-dependent proficiency increase in working memory. J Cogn Neurosci. 2008 Sep;20(9):1687–1697. doi: 10.1162/jocn.2008.20112. [DOI] [PubMed] [Google Scholar]
  • 9.Largus S. De compositionibus medicamentorum. Paris 1529. [Google Scholar]
  • 10.Marshall L, Molle M, Siebner HR, Born J. Bifrontal transcranial direct current stimulation slows reaction time in a working memory task. BMC Neurosci. 2005 Apr 8;6(1):23. doi: 10.1186/1471-2202-6-23. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Kellaway P. The part played by the electric fish in the early history of bioelectricity and electrotherapy. Bull Hist Med. 1946;20:112–137. [PubMed] [Google Scholar]
  • 12.Brunoni AR, Ferrucci R, Bortolomasi M, Vergari M, Tadini L, Boggio PS, et al. Transcranial direct current stimulation (tDCS) in unipolar vs. bipolar depressive disorder. Progress in neuro-psychopharmacology & biological psychiatry. 2011 Sep 17;35(1):96–101. doi: 10.1016/j.pnpbp.2010.09.010. [DOI] [PubMed] [Google Scholar]
  • 13.Boggio PS, Fregni F, Valasek C, Ellwood S, Chi R, Gallate J, et al. Temporal lobe cortical electrical stimulation during the encoding and retrieval phase reduces false memories. PLoS One. 2009;4(3):e4959. doi: 10.1371/journal.pone.0004959. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Zago S, Ferrucci R, Fregni F, Priori A. Bartholow, Sciamanna, Alberti: Pioneers in the Electrical Stimulation of the Exposed Human Cerebral Cortex. Neuroscientist. 2008 Jan 24;14(5):521–528. doi: 10.1177/1073858407311101. [DOI] [PubMed] [Google Scholar]
  • 15.Marshall L, Molle M, Hallschmid M, Born J. Transcranial direct current stimulation during sleep improves declarative memory. J Neurosci. 2004 Nov 3;24(44):9985–9992. doi: 10.1523/JNEUROSCI.2725-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Fecteau S, Knoch D, Fregni F, Sultani N, Boggio P, Pascual-Leone A. Diminishing risk-taking behavior by modulating activity in the prefrontal cortex: a direct current stimulation study. J Neurosci. 2007 Nov 14;27(46):12500–12505. doi: 10.1523/JNEUROSCI.3283-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Hecht D, Walsh V, Lavidor M. Transcranial direct current stimulation facilitates decision making in a probabilistic guessing task. J Neurosci. 2010 Mar 24;30(12):4241–4245. doi: 10.1523/JNEUROSCI.2924-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Brunoni AR, Valiengo L, Baccaro A, Zanao T, de Oliveira JF, Vieira GP, et al. Sertraline vs. electrical current therapy for treating depression clinical study - design, rationale and objectives. Contemp Clin Trials. Sep 17; doi: 10.1016/j.cct.2010.09.007. [DOI] [PubMed] [Google Scholar]
  • 19.Aldini G. Essai theorique et experimental sur le galvanisme. Paris 1804. [Google Scholar]
  • 20.Arndt R. Die electricitat in der psychiatrie. Arch Psychiat Nervenkrankh. 1869 Feb;:259–261. [Google Scholar]
  • 21.Kincses TZ, Antal A, Nitsche MA, Bartfai O, Paulus W. Facilitation of probabilistic classification learning by transcranial direct current stimulation of the prefrontal cortex in the human. Neuropsychologia. 2004;42(1):113–117. doi: 10.1016/s0028-3932(03)00124-6. [DOI] [PubMed] [Google Scholar]
  • 22.Stone DB, Tesche CD. Transcranial direct current stimulation modulates shifts in global/local attention. Neuroreport. 2009 Aug 5;20(12):1115–1119. doi: 10.1097/WNR.0b013e32832e9aa2. [DOI] [PubMed] [Google Scholar]
  • 23.Sparing R, Thimm M, Hesse MD, Kust J, Karbe H, Fink GR. Bidirectional alterations of interhemispheric parietal balance by non-invasive cortical stimulation. Brain. 2009 Jun 15;132(Pt 11):3011–3020. doi: 10.1093/brain/awp154. [DOI] [PubMed] [Google Scholar]
  • 24.Bolognini N, Fregni F, Casati C, Olgiati E, Vallar G. Brain polarization of parietal cortex augments training-induced improvement of visual exploratory and attentional skills. Brain research. 2010 Aug 19;1349:76–89. doi: 10.1016/j.brainres.2010.06.053. [DOI] [PubMed] [Google Scholar]
  • 25.Fertonani A, Rosini S, Cotelli M, Rossini PM, Miniussi C. Naming facilitation induced by transcranial direct current stimulation. Behav Brain Res. 2010 Apr 2;208(2):311–318. doi: 10.1016/j.bbr.2009.10.030. [DOI] [PubMed] [Google Scholar]
  • 26.Sparing R, Dafotakis M, Meister IG, Thirugnanasambandam N, Fink GR. Enhancing language performance with non-invasive brain stimulation--a transcranial direct current stimulation study in healthy humans. Neuropsychologia. 2008 Jan 15;46(1):261–268. doi: 10.1016/j.neuropsychologia.2007.07.009. [DOI] [PubMed] [Google Scholar]
  • 27.Priori A, Berardelli A, Rona S, Accornero N, Manfredi M. Polarization of the human motor cortex through the scalp. Neuroreport. 1998 Jul 13;9(10):2257–2260. doi: 10.1097/00001756-199807130-00020. [DOI] [PubMed] [Google Scholar]
  • 28.Nitsche MA, Paulus W. Excitability changes induced in the human motor cortex by weak transcranial direct current stimulation. J Physiol. 2000 Sep 15;527(Pt 3):633–639. doi: 10.1111/j.1469-7793.2000.t01-1-00633.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Fritsch B, Reis J, Martinowich K, Schambra HM, Ji Y, Cohen LG, et al. Direct current stimulation promotes BDNF-dependent synaptic plasticity: potential implications for motor learning. Neuron. 2010 Apr 29;66(2):198–204. doi: 10.1016/j.neuron.2010.03.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Nitsche MA, Liebetanz D, Antal A, Lang N, Tergau F, Paulus W. Modulation of cortical excitability by weak direct current stimulation--technical, safety and functional aspects. Suppl Clin Neurophysiol. 2003;56:255–276. doi: 10.1016/s1567-424x(09)70230-2. [DOI] [PubMed] [Google Scholar]
  • 31.Nitsche MA, Seeber A, Frommann K, Klein CC, Rochford C, Nitsche MS, et al. Modulating parameters of excitability during and after transcranial direct current stimulation of the human motor cortex. J Physiol. 2005 Oct 1;568(Pt 1):291–303. doi: 10.1113/jphysiol.2005.092429. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Kuo MF, Paulus W, Nitsche MA. Boosting focally-induced brain plasticity by dopamine. Cereb Cortex. 2008 Mar;18(3):648–651. doi: 10.1093/cercor/bhm098. [DOI] [PubMed] [Google Scholar]
  • 33.Wagner T, Fregni F, Fecteau S, Grodzinsky A, Zahn M, Pascual-Leone A. Transcranial direct current stimulation: a computer-based human model study. Neuroimage. 2007 Apr 15;35(3):1113–1124. doi: 10.1016/j.neuroimage.2007.01.027. [DOI] [PubMed] [Google Scholar]
  • 34.Miranda PC, Lomarev M, Hallett M. Modeling the current distribution during transcranial direct current stimulation. Clin Neurophysiol. 2006 Jul;117(7):1623–1629. doi: 10.1016/j.clinph.2006.04.009. [DOI] [PubMed] [Google Scholar]
  • 35.Bikson M, Datta A, Elwassif M. Establishing safety limits for transcranial direct current stimulation. Clin Neurophysiol. 2009 Jun;120(6):1033–1034. doi: 10.1016/j.clinph.2009.03.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Datta A, Bikson M, Fregni F. Transcranial direct current stimulation in patients with skull defects and skull plates: High-resolution computational FEM study of factors altering cortical current flow. Neuroimage. 2010 May 6;52(4):1268–1278. doi: 10.1016/j.neuroimage.2010.04.252. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Boggio PS, Alonso-Alonso M, Mansur CG, Rigonatti SP, Schlaug G, Pascual-Leone A, et al. Hand function improvement with low-frequency repetitive transcranial magnetic stimulation of the unaffected hemisphere in a severe case of stroke. Am J Phys Med Rehabil. 2006 Nov;85(11):927–930. doi: 10.1097/01.phm.0000242635.88129.38. [DOI] [PubMed] [Google Scholar]
  • 38.Iyer MB, Mattu U, Grafman J, Lomarev M, Sato S, Wassermann EM. Safety and cognitive effect of frontal DC brain polarization in healthy individuals. Neurology. 2005 Mar 8;64(5):872–875. doi: 10.1212/01.WNL.0000152986.07469.E9. [DOI] [PubMed] [Google Scholar]
  • 39.Fecteau S, Pascual-Leone A, Zald DH, Liguori P, Theoret H, Boggio PS, et al. Activation of prefrontal cortex by transcranial direct current stimulation reduces appetite for risk during ambiguous decision making. J Neurosci. 2007 Jun 6;27(23):6212–6218. doi: 10.1523/JNEUROSCI.0314-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.George MS, Aston-Jones G. Noninvasive techniques for probing neurocircuitry and treating illness: vagus nerve stimulation (VNS), transcranial magnetic stimulation (TMS) and transcranial direct current stimulation (tDCS) Neuropsychopharmacology. 2010 Jan;35(1):301–316. doi: 10.1038/npp.2009.87. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Fregni F, Pascual-Leone A. Technology insight: noninvasive brain stimulation in neurology-perspectives on the therapeutic potential of rTMS and tDCS. Nat Clin Pract Neurol. 2007 Jul;3(7):383–393. doi: 10.1038/ncpneuro0530. [DOI] [PubMed] [Google Scholar]
  • 42.Nitsche MA, Boggio PS, Fregni F, Pascual-Leone A. Treatment of depression with transcranial direct current stimulation (tDCS): a review. Exp Neurol. 2009 Sep;219(1):14–19. doi: 10.1016/j.expneurol.2009.03.038. [DOI] [PubMed] [Google Scholar]
  • 43.Brunoni AR, Tadini L, Fregni F. Changes in clinical trials methodology over time: a systematic review of six decades of research in psychopharmacology. PLoS One. 2010;5(3):e9479. doi: 10.1371/journal.pone.0009479. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Zhang X, Liu K, Sun J, Zheng Z. Safety and feasibility of repetitive transcranial magnetic stimulation (rTMS) as a treatment for major depression during pregnancy. Arch Womens Ment Health. 2010 Aug;13(4):369–370. doi: 10.1007/s00737-010-0163-5. [DOI] [PubMed] [Google Scholar]
  • 45.Bolognini N, Pascual-Leone A, Fregni F. Using non-invasive brain stimulation to augment motor training-induced plasticity. J Neuroeng Rehabil. 2009;6:8. doi: 10.1186/1743-0003-6-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Zaghi S, Acar M, Hultgren B, Boggio PS, Fregni F. Noninvasive Brain Stimulation with Low-Intensity Electrical Currents: Putative Mechanisms of Action for Direct and Alternating Current Stimulation. Neuroscientist. 2009 Dec;:29. doi: 10.1177/1073858409336227. [DOI] [PubMed] [Google Scholar]
  • 47.Nitsche MA, Cohen LG, Wassermann EM, Priori A, Lang N, Antal A, et al. Transcranial direct current stimulation: State of the art 2008 BRAIN STIMULATION. 2008;1(3):206–223. doi: 10.1016/j.brs.2008.06.004. [DOI] [PubMed] [Google Scholar]
  • 48.Priori A, Hallett M, Rothwell JC. Repetitive transcranial magnetic stimulation or transcranial direct current stimulation? Brain Stimul. 2009 Oct;2(4):241–245. doi: 10.1016/j.brs.2009.02.004. [DOI] [PubMed] [Google Scholar]
  • 49.Bindman LJ, Lippold OC, Redfearn JW. Relation between the Size and Form of Potentials Evoked by Sensory Stimulation and the Background Electrical Activity in the Cerebral Cortex of the Rat. J Physiol. 1964 May;171:1–25. doi: 10.1113/jphysiol.1964.sp007358. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Purpura DP, McMurtry JG. Intracellular Activities and Evoked Potential Changes during Polarization of Motor Cortex. J Neurophysiol. 1965 Jan 28;:166–185. doi: 10.1152/jn.1965.28.1.166. [DOI] [PubMed] [Google Scholar]
  • 51.Creutzfeldt OD, Fromm GH, Kapp H. Influence of transcortical d-c currents on cortical neuronal activity. Exp Neurol. 1962 Jun 5;:436–452. doi: 10.1016/0014-4886(62)90056-0. [DOI] [PubMed] [Google Scholar]
  • 52.Bindman LJ, Lippold OC, Redfearn JW. The Action of Brief Polarizing Currents on the Cerebral Cortex of the Rat (1) During Current Flow and (2) in the Production of Long-Lasting after-Effects. J Physiol. 1964 Aug 172;:369–382. doi: 10.1113/jphysiol.1964.sp007425. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Jefferys JG. Influence of electric fields on the excitability of granule cells in guinea-pig hippocampal slices. The Journal of Physiology. 1981;319:143–152. doi: 10.1113/jphysiol.1981.sp013897. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Bikson M, Inoue M, Akiyama H, Deans JK, Fox JE, Miyakawa H, et al. Effects of uniform extracellular DC electric fields on excitability in rat hippocampal slices in vitro. The Journal of Physiology. 2004 May 15;557(Pt 1):175–190. doi: 10.1113/jphysiol.2003.055772. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Nitsche MA, Paulus W. Sustained excitability elevations induced by transcranial DC motor cortex stimulation in humans. Neurology. 2001 Nov 27;57(10):1899–1901. doi: 10.1212/wnl.57.10.1899. [DOI] [PubMed] [Google Scholar]
  • 56.Liebetanz D, Nitsche MA, Tergau F, Paulus W. Pharmacological approach to the mechanisms of transcranial DC-stimulation-induced after-effects of human motor cortex excitability. Brain. 2002 Oct;125(Pt 10):2238–2247. doi: 10.1093/brain/awf238. [DOI] [PubMed] [Google Scholar]
  • 57.Nitsche MA, Fricke K, Henschke U, Schlitterlau A, Liebetanz D, Lang N, et al. Pharmacological modulation of cortical excitability shifts induced by transcranial direct current stimulation in humans. J Physiol. 2003 Nov 15;553(Pt 1):293–301. doi: 10.1113/jphysiol.2003.049916. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Stagg CJ, Best JG, Stephenson MC, O'Shea J, Wylezinska M, Kincses ZT, et al. Polarity-sensitive modulation of cortical neurotransmitters by transcranial stimulation. J Neurosci. 2009 Apr 22;29(16):5202–5206. doi: 10.1523/JNEUROSCI.4432-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.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 Oct 15;568(Pt 2):653–663. doi: 10.1113/jphysiol.2005.088310. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Cogiamanian F, Vergari M, Pulecchi F, Marceglia S, Priori A. Effect of spinal transcutaneous direct current stimulation on somatosensory evoked potentials in humans. Clin Neurophysiol. 2008 Nov;119(11):2636–2640. doi: 10.1016/j.clinph.2008.07.249. [DOI] [PubMed] [Google Scholar]
  • 61.Rango M, Cogiamanian F, Marceglia S, Barberis B, Arighi A, Biondetti P, et al. Myoinositol content in the human brain is modified by transcranial direct current stimulation in a matter of minutes: a 1H-MRS study. Magn Reson Med. 2008 Oct;60(4):782–789. doi: 10.1002/mrm.21709. [DOI] [PubMed] [Google Scholar]
  • 62.Boros K, Poreisz C, Munchau A, Paulus W, Nitsche MA. Premotor transcranial direct current stimulation (tDCS) affects primary motor excitability in humans. Eur J Neurosci. 2008 Mar;27(5):1292–1300. doi: 10.1111/j.1460-9568.2008.06090.x. [DOI] [PubMed] [Google Scholar]
  • 63.Lang N, Siebner HR, Ward NS, Lee L, Nitsche MA, Paulus W, et al. How does transcranial DC stimulation of the primary motor cortex alter regional neuronal activity in the human brain? Eur J Neurosci. 2005 Jul;22(2):495–504. doi: 10.1111/j.1460-9568.2005.04233.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Parra LC, Bikson M. Model of the effect of extracellular fields on spike time coherence; Conf Proc IEEE Eng Med Biol Soc; 2004. Jun, pp. 4584–4587. [DOI] [PubMed] [Google Scholar]
  • 65.Deans JK, Powell AD, Jefferys JG. Sensitivity of coherent oscillations in rat hippocampus to AC electric fields. The Journal of Physiology. 2007 Sep 1;583(Pt 2):555–565. doi: 10.1113/jphysiol.2007.137711. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Frohlich F, McCormick DA. Endogenous electric fields may guide neocortical network activity. Neuron. 2010 Jul 15;67(1):129–143. doi: 10.1016/j.neuron.2010.06.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Accornero N, Li Voti P, La Riccia M, Gregori B. Visual evoked potentials modulation during direct current cortical polarization. Exp Brain Res. 2007 Apr;178(2):261–261. doi: 10.1007/s00221-006-0733-y. [DOI] [PubMed] [Google Scholar]
  • 68.Matsunaga K, Nitsche MA, Tsuji S, Rothwell JC. Effect of transcranial DC sensorimotor cortex stimulation on somatosensory evoked potentials in humans. Clin Neurophysiol. 2004 Feb;115(2):456–460. doi: 10.1016/s1388-2457(03)00362-6. [DOI] [PubMed] [Google Scholar]
  • 69.Antal A, Kincses TZ, Nitsche MA, Bartfai O, Paulus W. Excitability changes induced in the human primary visual cortex by transcranial direct current stimulation: direct electrophysiological evidence. Invest Ophthalmol Vis Sci. 2004 Feb;45(2):702–707. doi: 10.1167/iovs.03-0688. [DOI] [PubMed] [Google Scholar]
  • 70.Galea JM, Jayaram G, Ajagbe L, Celnik P. Modulation of cerebellar excitability by polarity-specific noninvasive direct current stimulation. J Neurosci. 2009 Jul 15;29(28):9115–9122. doi: 10.1523/JNEUROSCI.2184-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Winkler T, Hering P, Straube A. Spinal DC stimulation in humans modulates post-activation depression of the H-reflex depending on current polarity. Clin Neurophysiol. 2010 Jun;121(6):957–961. doi: 10.1016/j.clinph.2010.01.014. [DOI] [PubMed] [Google Scholar]
  • 72.Merzagora AC, Foffani G, Panyavin I, Mordillo-Mateos L, Aguilar J, Onaral B, et al. Prefrontal hemodynamic changes produced by anodal direct current stimulation. Neuroimage. 2010 Feb 1;49(3):2304–2310. doi: 10.1016/j.neuroimage.2009.10.044. [DOI] [PubMed] [Google Scholar]
  • 73.Nitsche MA, Jaussi W, Liebetanz D, Lang N, Tergau F, Paulus W. Consolidation of human motor cortical neuroplasticity by D-cycloserine. Neuropsychopharmacology. 2004 Aug;29(8):1573–1578. doi: 10.1038/sj.npp.1300517. [DOI] [PubMed] [Google Scholar]
  • 74.Nitsche MA, Liebetanz D, Schlitterlau A, Henschke U, Fricke K, Frommann K, et al. GABAergic modulation of DC stimulation-induced motor cortex excitability shifts in humans. The European journal of neuroscience. 2004 May;19(10):2720–2726. doi: 10.1111/j.0953-816X.2004.03398.x. [DOI] [PubMed] [Google Scholar]
  • 75.Nitsche MA, Grundey J, Liebetanz D, Lang N, Tergau F, Paulus W. Catecholaminergic consolidation of motor cortical neuroplasticity in humans. Cereb Cortex. 2004 Nov;14(14):1240–1245. doi: 10.1093/cercor/bhh085. [DOI] [PubMed] [Google Scholar]
  • 76.Kuo MF, Unger M, Liebetanz D, Lang N, Tergau F, Paulus W, et al. Limited impact of homeostatic plasticity on motor learning in humans. Neuropsychologia. 2008;46(8):2122–2128. doi: 10.1016/j.neuropsychologia.2008.02.023. [DOI] [PubMed] [Google Scholar]
  • 77.Nitsche MA, Doemkes S, Karakose T, Antal A, Liebetanz D, Lang N, et al. Shaping the effects of transcranial direct current stimulation of the human motor cortex. J Neurophysiol. 2007 Apr;97(4):3109–3117. doi: 10.1152/jn.01312.2006. [DOI] [PubMed] [Google Scholar]
  • 78.Monte-Silva K, Kuo MF, Thirugnanasambandam N, Liebetanz D, Paulus W, Nitsche MA. Dose-dependent inverted U-shaped effect of dopamine (D2-like) receptor activation on focal and nonfocal plasticity in humans. J Neurosci. 2009 May 13;29(19):6124–6131. doi: 10.1523/JNEUROSCI.0728-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Nitsche MA, Kuo MF, Karrasch R, Wachter B, Liebetanz D, Paulus W. Serotonin affects transcranial direct current-induced neuroplasticity in humans. Biological psychiatry. 2009 Sep 1;66(5):503–508. doi: 10.1016/j.biopsych.2009.03.022. [DOI] [PubMed] [Google Scholar]
  • 80.Antal A, Kincses TZ, Nitsche MA, Paulus W. Manipulation of phosphene thresholds by transcranial direct current stimulation in man. Exp Brain Res. 2003 Jun;150(3):375–378. doi: 10.1007/s00221-003-1459-8. [DOI] [PubMed] [Google Scholar]
  • 81.Utz KS, Dimova V, Oppenlander K, Kerkhoff G. Electrified minds: Transcranial direct current stimulation (tDCS) and Galvanic Vestibular Stimulation (GVS) as methods of non-invasive brain stimulation in neuropsychology-A review of current data and future implications. Neuropsychologia. 2010 Jun 11;48(10):2789–2810. doi: 10.1016/j.neuropsychologia.2010.06.002. [DOI] [PubMed] [Google Scholar]
  • 82.Williams JA, Pascual-Leone A, Fregni F. Interhemispheric modulation induced by cortical stimulation and motor training. Phys Ther. 2010 Mar;90(3):398–410. doi: 10.2522/ptj.20090075. [DOI] [PubMed] [Google Scholar]
  • 83.Mahmoudi H, Haghighi AB, Petramfar P, Jahanshahi S, Salehi Z, Fregni F. Transcranial direct current stimulation: Electrode montage in stroke. Disability and Rehabilitation. doi: 10.3109/09638288.2010.532283. in press. [DOI] [PubMed] [Google Scholar]
  • 84.Redfearn JW, Lippold OC, Costain R. A Preliminary Account of the Clinical Effects of Polarizing the Brain in Certain Psychiatric Disorders. Br J Psychiatry. 1964 Nov;110:773–785. doi: 10.1192/bjp.110.469.773. [DOI] [PubMed] [Google Scholar]
  • 85.Cogiamanian F, Vergari M, Ardolino G, Scelzo E, Ciocca M, Barbieri M, et al., editors. Effects of transcutaneous spinal cord direct current stimulation (tsDCS) on the lower limb nociceptive flexion reflex in humans. Neurological Sciences; Proceedings of the XLI Congress of the Italian Neurological Society.2010. [Google Scholar]
  • 86.Datta A, Rahman A, Scaturro J, Bikson M. Electrode montages for tDCS and weak transcranial electrical stimulation Role of "return" electrode's position and size. Clin Neurophysiol. doi: 10.1016/j.clinph.2010.05.020. in press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Moliadze V, Antal A, Paulus W. Electrode-distance dependent after-effects of transcranial direct and random noise stimulation with extracephalic reference electrodes. Clinical Neurophysiology. 2010 Jun 14;121(12):2165–2171. doi: 10.1016/j.clinph.2010.04.033. [DOI] [PubMed] [Google Scholar]
  • 88.Datta A, Elwassif M, Battaglia F, Bikson M. Transcranial current stimulation focality using disc and ring electrode configurations: FEM analysis. J Neural Eng. 2008 Jun;5(2):163–174. doi: 10.1088/1741-2560/5/2/007. [DOI] [PubMed] [Google Scholar]
  • 89.Datta A, Bansal V, Diaz J, Patel J, Reato D, Bikson M. Gyri-precise head model of transcranial direct current stimulation: Improved spatial focality using a ring electrode versus conventional rectangular pad. Brain Stimulation. 2009;2(4):201–207. doi: 10.1016/j.brs.2009.03.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Sadleir RJ, Vannorsdall TD, Schretlen DJ, Gordon B. Transcranial direct current stimulation (tDCS) in a realistic head model. Neuroimage. 2010 Jul 15;51(4):1310–1318. doi: 10.1016/j.neuroimage.2010.03.052. [DOI] [PubMed] [Google Scholar]
  • 91.Minhas P, Bansal V, Patel J, Ho JS, Diaz J, Datta A, et al. Electrodes for high-definition transcutaneous DC stimulation for applications in drug delivery and electrotherapy, including tDCS. Journal of Neuroscience Methods. 2010 Jul 15;190(2):188–197. doi: 10.1016/j.jneumeth.2010.05.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Faria P, Leal A, Miranda PC. Comparing different electrode configurations using the 10-10 international system in tDCS: a finite element model analysis. Conf Proc IEEE Eng Med Biol Soc. 2009;2009:1596–1599. doi: 10.1109/IEMBS.2009.5334121. [DOI] [PubMed] [Google Scholar]
  • 93.Suh HS, Kim SH, Lee WH, Kim TS. Realistic simulation of transcranial direct current stimulation via 3-d high-resolution finite element analysis: Effect of tissue anisotropy. Conf Proc IEEE Eng Med Biol Soc. 2009;2009:638–641. doi: 10.1109/IEMBS.2009.5333686. [DOI] [PubMed] [Google Scholar]
  • 94.Liebetanz D, Koch R, Mayenfels S, Konig F, Paulus W, Nitsche MA. Safety limits of cathodal transcranial direct current stimulation in rats. Clin Neurophysiol. 2009 Jun;120(6):1161–1167. doi: 10.1016/j.clinph.2009.01.022. [DOI] [PubMed] [Google Scholar]
  • 95.Fregni F, Liebetanz D, Monte-Silva KK, Oliveira MB, Santos AA, Nitsche MA, et al. Effects of transcranial direct current stimulation coupled with repetitive electrical stimulation on cortical spreading depression. Exp Neurol. 2007 Mar;204(1):462–466. doi: 10.1016/j.expneurol.2006.09.019. [DOI] [PubMed] [Google Scholar]
  • 96.Palm U, Keeser D, Schiller C, Fintescu Z, Reisinger E, Nitsche M, et al. Skin lesions after treatment with transcranial direct current stimulation (tDCS) Brain Stimul. 2008 Oct;1(4):386–387. doi: 10.1016/j.brs.2008.04.003. [DOI] [PubMed] [Google Scholar]
  • 97.Durand S, Fromy B, Bouye P, Saumet JL, Abraham P. Vasodilatation in response to repeated anodal current application in the human skin relies on aspirin-sensitive mechanisms. J Physiol. 2002 Apr 1;540(Pt 1):261–269. doi: 10.1113/jphysiol.2001.013364. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Poreisz C, Boros K, Antal A, Paulus W. Safety aspects of transcranial direct current stimulation concerning healthy subjects and patients. Brain Res Bull. 2007 May 30;72(4–6):208–214. doi: 10.1016/j.brainresbull.2007.01.004. [DOI] [PubMed] [Google Scholar]
  • 99.Tadini L, El-Nazer R, Brunoni AR, Williams J, Carvas M, Boggio PS, et al. Cognitive, mood and EEG effects of noninvasive cortical stimulation with weak electrical currents. The Journal of Electroconvulsive Therapy. 2010 doi: 10.1097/YCT.0b013e3181e631a8. [DOI] [PubMed] [Google Scholar]
  • 100.Nitsche MA, Liebetanz D, Lang N, Antal A, Tergau F, Paulus W. Safety criteria for transcranial direct current stimulation (tDCS) in humans. Clin Neurophysiol. 2003 Nov;114(11):2220–2222. doi: 10.1016/s1388-2457(03)00235-9. author reply 2–3. [DOI] [PubMed] [Google Scholar]
  • 101.Fregni F, Boggio PS, Nitsche MA, Rigonatti SP, Pascual-Leone A. Cognitive effects of repeated sessions of transcranial direct current stimulation in patients with depression. Depress Anxiety. 2006;23(8):482–484. doi: 10.1002/da.20201. [DOI] [PubMed] [Google Scholar]
  • 102.Boggio PS, Nunes A, Rigonatti SP, Nitsche MA, Pascual-Leone A, Fregni F. Repeated sessions of noninvasive brain DC stimulation is associated with motor function improvement in stroke patients. Restor Neurol Neurosci. 2007;25(2):123–129. [PubMed] [Google Scholar]
  • 103.Loo C, Martin D, Pigot M, Arul-Anandam P, Mitchell P, Sachdev P. Transcranial Direct Current Stimulation Priming of Therapeutic Repetitive Transcranial Magnetic Stimulation: A Pilot Study. J ECT. 2009 May 12; doi: 10.1097/YCT.0b013e3181a2f87e. [DOI] [PubMed] [Google Scholar]
  • 104.Fregni F, Thome-Souza S, Nitsche MA, Freedman SD, Valente KD, Pascual-Leone A. A controlled clinical trial of cathodal DC polarization in patients with refractory epilepsy. Epilepsia. 2006 Feb;47(2):335–342. doi: 10.1111/j.1528-1167.2006.00426.x. [DOI] [PubMed] [Google Scholar]
  • 105.Arul-Anandam AP, Loo C, Mitchell P. Induction of hypomanic episode with transcranial direct current stimulation. J ECT. 2010 Mar;26(1):68–69. doi: 10.1097/YCT.0b013e3181a744bf. [DOI] [PubMed] [Google Scholar]
  • 106.Baccaro A, Brunoni AR, Bensenor IM, Fregni F. Hypomanic episode in unipolar depression during transcranial direct current stimulation. Acta Neuropsychiatrica. 2010;22(6):316–318. [Google Scholar]
  • 107.Brunoni AR, Valiengo L, Zanao T, Oliveira J, Bensenor IM, Fregni F. Manic psychosis following transcranial direct current stimulation and sertraline. The Journal of Neuropsychiatry and Clinical Neurosciences. doi: 10.1176/jnp.23.3.jnpe4. in press. [DOI] [PubMed] [Google Scholar]
  • 108.Miranda PC, Faria P, Hallett M. What does the ratio of injected current to electrode area tell us about current density in the brain during tDCS? Clin Neurophysiol. 2009 Jun;120(6):1183–1187. doi: 10.1016/j.clinph.2009.03.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Fregni F, Boggio PS, Valle AC, Otachi P, Thut G, Rigonatti SP, et al. Homeostatic effects of plasma valproate levels on corticospinal excitability changes induced by 1Hz rTMS in patients with juvenile myoclonic epilepsy. Clin Neurophysiol. 2006 Jun;117(6):1217–1227. doi: 10.1016/j.clinph.2006.02.015. [DOI] [PubMed] [Google Scholar]
  • 110.Ziemann U. Pharmacology of TMS. Suppl Clin Neurophysiol. 2003;56:226–231. [PubMed] [Google Scholar]
  • 111.Bajwa S, Bermpohl F, Rigonatti SP, Pascual-Leone A, Boggio PS, Fregni F. Impaired interhemispheric interactions in patients with major depression. J Nerv Ment Dis. 2008 Sep;196(9):671–677. doi: 10.1097/NMD.0b013e318183f86f. [DOI] [PubMed] [Google Scholar]
  • 112.Soubasi E, Chroni E, Gourzis P, Zisis A, Beratis S, Papathanasopoulos P. Cortical motor neurophysiology of patients with schizophrenia: a study using transcranial magnetic stimulation. Psychiatry research. 2010 Apr 30;176(2–3):132–136. doi: 10.1016/j.psychres.2009.04.003. [DOI] [PubMed] [Google Scholar]
  • 113.Mhalla A, de Andrade DC, Baudic S, Perrot S, Bouhassira D. Alteration of cortical excitability in patients with fibromyalgia. Pain. 2010 Jun;149(3):495–500. doi: 10.1016/j.pain.2010.03.009. [DOI] [PubMed] [Google Scholar]
  • 114.Brighina F, Palermo A, Fierro B. Cortical inhibition and habituation to evoked potentials: relevance for pathophysiology of migraine. J Headache Pain. 2009 Apr;10(2):77–84. doi: 10.1007/s10194-008-0095-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Chaieb L, Antal A, Paulus W. Gender-specific modulation of short-term neuroplasticity in the visual cortex induced by transcranial direct current stimulation. Vis Neurosci. 2008 Jan-Feb;25(1):77–81. doi: 10.1017/S0952523808080097. [DOI] [PubMed] [Google Scholar]
  • 116.Ferri R, Del Gracco S, Elia M, Musumeci SA. Age-related changes of cortical excitability in subjects with sleep-enhanced centrotemporal spikes: a somatosensory evoked potential study. Clin Neurophysiol. 2000 Apr;111(4):591–599. doi: 10.1016/s1388-2457(99)00249-7. [DOI] [PubMed] [Google Scholar]
  • 117.Lang N, Hasan A, Sueske E, Paulus W, Nitsche MA. Cortical hypoexcitability in chronic smokers? A transcranial magnetic stimulation study. Neuropsychopharmacology. 2008 Sep;33(10):2517–2523. doi: 10.1038/sj.npp.1301645. [DOI] [PubMed] [Google Scholar]
  • 118.Monte-Silva K, Liebetanz D, Grundey J, Paulus W, Nitsche MA. Dosage-dependent non-linear effect of L-dopa on human motor cortex plasticity. J Physiol. 2010 Sep 15;588(Pt 18):3415–3424. doi: 10.1113/jphysiol.2010.190181. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Ohn SH, Park CI, Yoo WK, Ko MH, Choi KP, Kim GM, et al. Time-dependent effect of transcranial direct current stimulation on the enhancement of working memory. Neuroreport. 2008 Jan 8;19(1):43–47. doi: 10.1097/WNR.0b013e3282f2adfd. [DOI] [PubMed] [Google Scholar]
  • 120.Boggio PS, Ferrucci R, Rigonatti SP, Covre P, Nitsche M, Pascual-Leone A, et al. Effects of transcranial direct current stimulation on working memory in patients with Parkinson's disease. J Neurol Sci. 2006 Nov 1;249(1):31–38. doi: 10.1016/j.jns.2006.05.062. [DOI] [PubMed] [Google Scholar]
  • 121.Froc DJ, Chapman CA, Trepel C, Racine RJ. Long-term depression and depotentiation in the sensorimotor cortex of the freely moving rat. J Neurosci. 2000 Jan 1;20(1):438–445. doi: 10.1523/JNEUROSCI.20-01-00438.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Monte-Silva KK, Kuo MF, Liebetanz D, Paulus W, Nitsche MA. Shaping the optimal repetition interval for cathodal transcranial direct current stimulation (tDCS) J Neurophysiol. 2010 Jan 27;103(4):1735–1740. doi: 10.1152/jn.00924.2009. [DOI] [PubMed] [Google Scholar]
  • 123.Siebner HR, Lang N, Rizzo V, Nitsche MA, Paulus W, Lemon RN, et al. Preconditioning of low-frequency repetitive transcranial magnetic stimulation with transcranial direct current stimulation: evidence for homeostatic plasticity in the human motor cortex. J Neurosci. 2004 Mar 31;24(13):3379–3385. doi: 10.1523/JNEUROSCI.5316-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Brunoni AR, Fregni F. Clinical Trial Design in Noninvasive Brain Stimulation Psychiatric Research. International Journal of Methods in Psychiatry Research. doi: 10.1002/mpr.338. (in press) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Gandiga PC, Hummel FC, Cohen LG. Transcranial DC stimulation (tDCS): a tool for double-blind sham-controlled clinical studies in brain stimulation. Clin Neurophysiol. 2006 Apr;117(4):845–850. doi: 10.1016/j.clinph.2005.12.003. [DOI] [PubMed] [Google Scholar]
  • 126.Dundas JE, Thickbroom GW, Mastaglia FL. Perception of comfort during transcranial DC stimulation: effect of NaCl solution concentration applied to sponge electrodes. Clin Neurophysiol. 2007 May;118(5):1166–1170. doi: 10.1016/j.clinph.2007.01.010. [DOI] [PubMed] [Google Scholar]
  • 127.Freedman B. Equipoise and the ethics of clinical research. N Engl J Med. 1987 Jul 16;317(3):141–145. doi: 10.1056/NEJM198707163170304. [DOI] [PubMed] [Google Scholar]
  • 128.Schutter DJ. Nosce te ipsum: On the efficacy of transcranial magnetic stimulation in major depressive disorder. Biol Psychiatry. 2010 Mar 1;67(5):e27. doi: 10.1016/j.biopsych.2009.10.035. author reply e9. [DOI] [PubMed] [Google Scholar]
  • 129.Yu E, Lurie P. Transcranial magnetic stimulation not proven effective. Biol Psychiatry. 2010 Jan 15;67(2):e13. doi: 10.1016/j.biopsych.2009.03.026. author reply e5–7. [DOI] [PubMed] [Google Scholar]
  • 130.D'Agostino RB., Jr Debate: The slippery slope of surrogate outcomes. Curr Control Trials Cardiovasc Med. 2000;1(2):76–78. doi: 10.1186/cvm-1-2-076. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Kobayashi M, Pascual-Leone A. Transcranial magnetic stimulation in neurology. Lancet Neurology. 2003 Mar;2(3):145–156. doi: 10.1016/s1474-4422(03)00321-1. [DOI] [PubMed] [Google Scholar]
  • 132.Pascual-Marqui RD, Esslen M, Kochi K, Lehmann D. Functional imaging with low-resolution brain electromagnetic tomography (LORETA): a review. Methods Find Exp Clin Pharmacol. 2002;24(Suppl C):91–95. [PubMed] [Google Scholar]
  • 133.May A, Hajak G, Ganssbauer S, Steffens T, Langguth B, Kleinjung T, et al. Structural brain alterations following 5 days of intervention: dynamic aspects of neuroplasticity. Cerebral Cortex. 2007 Jan;17(1):205–210. doi: 10.1093/cercor/bhj138. [DOI] [PubMed] [Google Scholar]
  • 134.Kriegeskorte N, Lindquist MA, Nichols TE, Poldrack RA, Vul E. Everything you never wanted to know about circular analysis, but were afraid to ask. J Cereb Blood Flow Metab. 2010 Jun 23;30(9):1551–1557. doi: 10.1038/jcbfm.2010.86. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Gelenberg AJ, Thase ME, Meyer RE, Goodwin FK, Katz MM, Kraemer HC, et al. The history and current state of antidepressant clinical trial design: a call to action for proof-of-concept studies. J Clin Psychiatry. 2008 Oct;69(10):1513–1528. doi: 10.4088/jcp.v69n1001. [DOI] [PubMed] [Google Scholar]
  • 136.Brunoni AR, Lopes M, Fregni F. A systematic review and meta-analysis of clinical studies on major depression and BDNF levels: implications for the role of neuroplasticity in depression. Int J Neuropsychopharmacol. 2008 Dec;11(8):1169–1180. doi: 10.1017/S1461145708009309. [DOI] [PubMed] [Google Scholar]
  • 137.Dowlati Y, Herrmann N, Swardfager W, Liu H, Sham L, Reim EK, et al. A meta-analysis of cytokines in major depression. Biol Psychiatry. 2010 Mar 1;67(5):446–457. doi: 10.1016/j.biopsych.2009.09.033. [DOI] [PubMed] [Google Scholar]
  • 138.Quinones MP, Kaddurah-Daouk R. Metabolomics tools for identifying biomarkers for neuropsychiatric diseases. Neurobiol Dis. 2009 Aug;35(2):165–176. doi: 10.1016/j.nbd.2009.02.019. [DOI] [PubMed] [Google Scholar]
  • 139.Marques AH, Cizza G, Sternberg E. [Brain-immune interactions and implications in psychiatric disorders] Rev Bras Psiquiatr. 2007 May;29(Suppl 1):S27–S32. doi: 10.1590/s1516-44462007000500006. [DOI] [PubMed] [Google Scholar]
  • 140.Maes M, Yirmyia R, Noraberg J, Brene S, Hibbeln J, Perini G, et al. The inflammatory & neurodegenerative (I&ND) hypothesis of depression: leads for future research and new drug developments in depression. Metab Brain Dis. 2009 Mar;24(1):27–53. doi: 10.1007/s11011-008-9118-1. [DOI] [PubMed] [Google Scholar]
  • 141.Spencer-Smith M, Anderson V. Healthy and abnormal development of the prefrontal cortex. Dev Neurorehabil. 2009;12(5):279–297. doi: 10.3109/17518420903090701. [DOI] [PubMed] [Google Scholar]
  • 142.Frye RE, Rotenberg A, Ousley M, Pascual-Leone A. Transcranial magnetic stimulation in child neurology: current and future directions. J Child Neurol. 2008 Jan;23(1):79–96. doi: 10.1177/0883073807307972. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Fumagalli M, Vergari M, Pasqualetti P, Marceglia S, Mameli F, Ferrucci R, et al. Brain switches utilitarian behavior: does gender make the difference? PLoS One. 2010;5(1):e8865. doi: 10.1371/journal.pone.0008865. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Luber B, Fisher C, Appelbaum PS, Ploesser M, Lisanby SH. Non-invasive brain stimulation in the detection of deception: scientific challenges and ethical consequences. Behav Sci Law. 2009;27(2):191–208. doi: 10.1002/bsl.860. [DOI] [PubMed] [Google Scholar]
  • 145.Mameli F, Mrakic-Sposta S, Vergari M, Fumagalli M, Macis M, Ferrucci R, et al. Dorsolateral prefrontal cortex specifically processes general - but not personal - knowledge deception: Multiple brain networks for lying. Behav Brain Res. 2010 Aug 25;211(2):164–168. doi: 10.1016/j.bbr.2010.03.024. [DOI] [PubMed] [Google Scholar]
  • 146.Boggio PS, Campanha C, Valasek CA, Fecteau S, Pascual-Leone A, Fregni F. Modulation of decision-making in a gambling task in older adults with transcranial direct current stimulation. Eur J Neurosci. 2010 Feb;31(3):593–597. doi: 10.1111/j.1460-9568.2010.07080.x. [DOI] [PubMed] [Google Scholar]
  • 147.Reis J, Robertson E, Krakauer JW, Rothwell J, Marshall L, Gerloff C, et al. Consensus: "Can tDCS and TMS enhance motor learning and memory formation?". Brain Stimul. 2008 Oct;1(4):363–369. doi: 10.1016/j.brs.2008.08.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Anonymous. [cited 2010 September 25th];2010 Available from: http://www.youtube.com/verify_age?next_url=http://www.youtube.com/watch%3Fv%3DSymQqZYiVyA.

RESOURCES