Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2019 Dec 1.
Published in final edited form as: Curr Opin Biomed Eng. 2018 Sep 28;8:38–44. doi: 10.1016/j.cobme.2018.09.006

Physiological effects of low-magnitude electric fields on brain activity: advances from in vitro, in vivo and in silico models

Julien Modolo 1, Yves Denoyer 1, Fabrice Wendling 1,*, Pascal Benquet 1
PMCID: PMC6516771  NIHMSID: NIHMS1508428  PMID: 31106284

Abstract

While electrical stimulation of brain tissue has been thoroughly investigated over the last decades, ongoing questions remain regarding the neurophysiological effects of low-level electric fields (on the order of 1 V/m) on brain activity. Electric fields at such levels are, for example, induced by transcranial direct/alternating current stimulation (tDCS/tACS). Action potentials can be indeed elicited when applied (supra-threshold) electric fields are in the 10–100 V/m range, while lower (subthreshold) electric fields result in more limited and subtler membrane polarization effects. In this review, we address the question of the mechanisms underlying the immediate effects (also referred to as acute, concurrent or short-term) and the lasting effects (also referred to as long-term or aftereffects) of low-level electric fields on brain tissue. We review recent evidence at the in vitro and in vivo (animal and human) level, and also present mechanistic insights gained from in silico models, which are still few but have received increased attention over the recent past years. We highlight the convergent evidence towards potential mechanisms, and also discuss discrepancies between in vitro studies and human tDCS/tACS studies that require further investigation to bridge the gap between the single-cell and large-scale network level. Possible novel avenues of research are discussed.

Keywords: Electric fields, transcranial direct current stimulation (tDCS), transcranial alternating current stimulation (tACS), electrophysiology, in vitro, in vivo, computational, model

1. Introduction

Electrical stimulation has established as a standard clinical practice in a number of clinical applications, ranging rom diagnostic (e.g., functional stimulation during stereoelectroencephalography -SEEG- in drug-refractory epileptic patients) to therapeutic (e.g., chronic stimulation of deep brain regions in Parkinson’s disease to alleviate motror symptoms) procedures. A common point of these applications of electrical stimulation is that they are based on the use of suprathreshold electric fields, i.e. electric fields that are sufficiently strong to trigger action potentials at the level of single cells. However, increasingly popular neuromodulation techniques such as transcranial current stimulation (tCS) are based on the use of electric currents (typically up to 2 mA), either direct (tDCS) or alternating (tACS), delivered at the level of the scalp. Effects of transcranial electrical stimulation depend on the surface area of the stimulating electrodes. Since current density corresponds to the (stimulation strength (A)/electrode size (cm2)) ratio, tDCS electrodes/wet sponges used in human with a nominal size of 25–35 cm2 and currents of 2 mA result in current densities on the order of ~0.01–0.05 mA/cm2 (Nitsche et al., 2003; Thair et al., 2017). The fraction of current reaching the cortex also depends on the electrode placement as shown by pioneer work of (Rush and Driscoll 1968). For instance, for bipolar electrodes positioned on the centers of the occipital and frontal bones, the total current passing through the cranial cavity was about 45% of that applied. For an electrode spacing less than 5 cm, the shunting effect of the skull and cerebrospinal fluid -CSF- predominates. tCS exerts multiple effects on brain tissue activity, with detectable effects at the behavioral and/or clinical level. While conflicting data exists in the literature on the consistency of these effects, there is now converging evidence that the electric fields induced by tDCS/tACS, even if subthreshold (on the order of 1 V/m, Fig 1A), can modulate neuronal function. In particular, it is estimated that a tDCS intensity of 1 mA generates electrical fields at the level of cortical neurons in the 0.2–0.5 V/m range (Miranda, Lomarev et al. 2006, Datta, Bansal et al. 2009, Sadleir, Vannorsdall et al. 2010), while 2 mA (~0.01mA/cm2) would produce an electric field of about of 1 V/m (Jackson, Rahman et al. 2016, Voroslakos, Takeuchi et al. 2018). Interestingly, for those values (tDCS at an intensity of 2 mA generating an electric field of 1 V/m), the voltage gradient on the neuronal resting membrane potential is estimated on the order of 0.2 mV (Bikson, Inoue et al. 2004), a value far from being able to bring neurons to spiking threshold.

Figure 1.

Figure 1

A: Effects of transcranial low-level electric fields depends on cortical neuron orientation and brain geometry, but 1 mA generates electrical fields < 1 V/m and induce very limited polarization effect on the resting membrane potential. The right panel illustrates the orientation of cortical neurons apical dendrites perpendicular to the gyri surface (photography from the cortex of GFP-transgenic mice) and superimposed to the theoretical electric field.

B: Computational illustration of the network-wide amplification of weak (sub-threshold) field effects under anodal tDCS.

B1: A detailed computational model (unpublished) comprising 100 000 synapses connecting 1000 neurons (principal neurons and GABAergic interneurons (i.e. basket, somatostatin-positive -SST-, vasoactive intestinal polypeptide-positive -VIP-). tDCS modulates neurotransmission by slightly modifying the probability of release and post synaptic potential of each individual synapse, depending on its position in the electric field.

B2: In the model presented in (B1), simulated tDCS does not affect the neurophysiological properties of an isolated cortical pyramidal cell. This illustration shows that the firing pattern of an individual neuron does not change without (top) or in the presence (bottom) of simulated tDCS effects.

B3: In the same configuration than (B2) in the absence of visible effects on a single neuron, tDCS induces a network modulation. Left. Superimposed firing patterns of 1000 neurons are shown (red traces: pyramidal neurons; green: basket cells; blue : SSt interneurons; yellow : VIP Interneurons) in absence (top) or presence (bottom) of simulated tDCS. Intra = simulated intracellular recording of individual neurons. The corresponding simulated local field potential (LFP) is depicted in gray upon individual traces. Note that, in the presence of simulated tDCS, the synchronization pattern of the neural network LFP changes as shown in the respective power spectrum (right). a.u: arbitrary unit. PSD: Power spectrum density.

C. Illustration of transcranially-induced weak fields slightly polarizing axon terminals and modulating synaptic transmission. Note that presynaptic synapses compartments are not uniformly oriented. Transcranial current stimulation might also affect glial cells such as astrocytes (grey).

D. Synthetic view on the effects of exogenous and endogenous fields. A positive electric field is oriented along decreasing potentials (grey arrow, positive EEXO>0). A tDCS-range exogenous field has minor effects on neuronal activity and excitability. In the opposite direction (red arrow, negative field EEXO<0), the field may slightly depolarize the soma (close to the anode) and hyperpolarize the apical dendrite due to induced currents flowing across the transmembrane resistance. Excitability is increased in this case. In addition, impacted neurons create themselves an endogenous field (blue arrow, EENDO ) that will also influence passive neighbouring neurons (illustrated on the right as a passive neuron). Such possible ephaptic interactions possibly contribute to changes in excitability and synchronization at the network level. VTM=transmembrane membrane potential. I = current. Adapted from Brain Stimulation: models, experiments and open questions; DELIVERABLE D1.1: Review of state-of-the-art in currents distribution and effects, Hive Project, EU FP7 FET OPEN – 222079, 2009.

Therefore, the main question that arises is: what are the mechanisms that explain neurophysiological effects, for such levels of in situ electric field? Among the reported effects of induced electric fields in humans, the most reproducible effect at the lowest known electric field levels (< 0.1 V/m in situ at 20 Hz, (IEEE 2010, International Commission on Non-Ionizing Radiation 2010, Hirata, Takano et al. 2011, Laakso and Hirata 2012)) is the perception of electro- and magneto-phosphenes, consisting in the perception of flickering lights due to the currents induced in the brain, and most certainly within the retina (Lovsund, Oberg et al. 1980, Attwell 2003). However, the retina has a very specific architecture adapted for the amplification of weak signals (e.g., convergent pathways between photoreceptors and bipolar cells (Attwell 2003)), which cannot directly be translated to the rest of the brain. A considerable number of studies have therefore attempted to investigate the question of interaction mechanisms of electric fields in the 1 V/m range, from cultures of cells (in vitro) to animal experiments (in vivo). Significant efforts have also been done in the field of computational modeling (in silico) to integrate these experimental results. Therefore, in the following, we focus on the main results on the exploration of interaction mechanisms, approached with experimental and modeling techniques. Furthermore, for each level of description, we mention the results that have been obtained both in terms of acute and lasting effects. By “acute”, we refer to immediate effects occurring concurrently with tCS (i.e during stimulation) or on the short term (i.e seconds after stimulation) while the terminology “lasting” refers to after-effects observed on a longer term (i.e. hours or days after stimulation).

2. In vitro approaches

Most in vitro studies have been performed using an experimental setup consisting of two long parallel wires or plates placed in the bath where hippocampal or cortical slices are maintained. The major advantages of this specific setup are the generation of a quasi-uniform electric field across the entire tissue, and the possibility to measure (e.g., patch clamp) the impact on cellular properties for different DC stimulation parameters. In this setup, the effects on neurophysiology are strongly dependent on the electric field orientation and stimulation intensity (Fig 1A), as mentioned in (Kabakov, Muller et al. 2012) and (Rahman, Reato et al. 2013). Among the numerous effects reported in vitro, a number of significant modifications have been reported: in calcium entry, depolarization, hyperpolarization, firing rate, spike timing, neuronal input/output function, modulation of synaptic transmission, modulation of NMDA-dependent long-term potentiation, modulation of mGluRdependent long-term depression (Sun, Lipton et al. 2016) and even morphological changes such as axonal growth and guidance (Jackson, Rahman et al. 2016, Kronberg, Bridi et al. 2017, Lafon, Rahman et al. 2017).

However, one major drawback of most DC stimulation in vitro studies is that they involve electric field magnitudes of 20 to 60 V/m, which is at least 20 times higher than the electric field induced in the human brain for a 2mA tDCS (~1 V/m). It is worth noting though that no proper modeling of the whole experimental setup (slice and recording chamber filled in with ACSF), including electric fields induced in the depth of the slice, has been reported so far and actual values are likely lower. Therefore, the direct correspondence between these in vitro results and the neurophysiological mechanisms triggered by non-invasive electrical brain stimulation (which involves much weaker in situ fields) is not obvious. Furthermore, one should mention that, in the highly folded human brain cortex, tDCS-induced electric fields generated by two or more distant electrodes is not uniform. This limitation was addressed in a few in vitro studies which made use of more realistic tDCS/tACS-level electric fields (0.5 to 1 V/m) stimulation. Results confirmed an average of 0.12 mV depolarization of the cell bodies for each 1 V/m of applied field (Jefferys, Deans et al. 2003), and a modulation of axon terminals polarization that are two–three fold more sensitive than somas (Reato, Rahman et al. 2010). While the impact of electric fields of such magnitude on single neuron activity is almost negligible, the situation might be different at the level of local neural networks involving several thousand neurons. One of the possible mechanisms of tDCS might indeed be the network-wide amplification of weak membrane depolarizations (Frohlich and McCormick 2010), as illustrated by computational simulations in Fig 1B. Since tDCS impacts mainly axon terminals, and since neurotransmitter release is very sensitive to presynaptic bouton polarization (Debanne, Campanac et al. 2011), the main effect of tDCS might actually be on synaptic transmission (Fig 1C) and presynaptic modulation of short-term plasticity (Rahman, Lafon et al. 2017), even if such mechanisms remain to be demonstrated for an electric field on the order of 1 V/m. Finally, one drawback of most of these experiments is the lack of a corresponding computational model that would account for several factors, among which the relative orientation between the electric field and neurons, the differential effect of the electric field on specific locations of neuronal complex geometry. Therefore, conclusions from these studies have to be taken with caution, since the interplay between electric fields and neuronal type and geometry can greatly affect the outcome.

3. In vivo approaches

In vivo experiments can highlight some after-effects of tDCS on longer time scales (days) as compared to in vitro studies (hours). As aforementioned, network-based mechanisms might be one fundamental mechanism of tDCS-level electric fields. Therefore, it is required to analyze the response of brain tissue with intact connectivity to such fields, which has been the motivation in a number of recent in vivo studies, mostly conducted in rodents. It has been shown that transcranial current anodal stimulation in mice could facilitate memory (preference towards novel objects), an effect that persisted one week after stimulation and likely explained by increasing hippocampal long-term plasticity.

Glutamatergic circuits seems to be particularly affected by direct current stimulation. NMDAreceptor dependent long-term synaptic potentiation (LTP) (Fritsch, Reis et al. 2010) and AMPA receptor translocation and phosphorylation (Stafford, Brownlow et al. 2018) can be induced by direct current stimulation in vivo.

Interestingly, blockade of the TrkB brain-derived neurotrophic factor (BDNF) receptor reduced tDCS facilitating effects on memory and tDCS-induced long-lasting changes in BDNF expression by epigenetic mechanisms (Podda, Cocco et al. 2016), suggesting that tDCS might have, at least indirectly, effects on BDNF regulation. Inversely, long-term depression (LTD) evoked in the somatosensory cortex behaving rabbit after cathodal tDCS is prevented by adenosine A1 receptors blockade, suggesting that tDCS might influence adenosine release in glutamatergic circuits (MarquezRuiz, Leal-Campanario et al. 2012).

Another possible mechanism of tDCS is the indirect neuromodulation of cortical circuits through astrocytes. Even if this glial cell type is not excitable per se, and the membrane potential is stable, it seems sensitive to electric fields. Among the reported effects, increased glucose metabolism was found in cultured cortical mouse astrocytes under low magnitude electric fields (0.3 V/m) induced by tDCS (Huang, Peng et al. 1997). Recent studies using the calcium indicator G-CAMP7 expression in mice astrocytes in vivo, pointed at the indirect activation of astrocytic networks by tDCS, through the release of neurotransmitters such as noradrenaline (Monai, Ohkura et al. 2016, Monai and Hirase 2018). Since the activity of glial cells (astrocyte, microglia) interferes with synaptic plasticity, their modulation provides a possible explanation to the well-established lasting effect of tDCS (Gellner, Reis et al. 2016). Interestingly, based on the predictions of a computational model, a recent study (Mina, Modolo et al. 2017), tested experimentally the possibility to modulate epileptic hippocampal paroxysmal discharges (HPDs), which are observed in a mouse model of epilepsy, using DC stimulation delivered intracranially. A drastic reduction of HPDs was observed when anodal stimulation was delivered, generating an electric field on average < 5 V/m in the hippocampus. More specifically, in terms of underlying mechanisms, the model also predicted that the effect was mainly mediated by the depolarization of slow, dendritic-targeting GABAergic inhibitory interneurons, a testable hypothesis.

4. In silico approaches

As illustrated by the two previous sections, and despite the many in vitro and in vivo studies which have been conducted to investigate the behavior of small and large networks of neurons under the effect of electric or magnetic stimulations, the mechanisms by which low-level E-fields (i.e. compatible with those induced by tCS) impact the activity of neuronal assemblies remain elusive, as is the link between immediate and after-effects. In this context, in silico models also constitute a promising - though probably more challenging - approach to quantitatively analyze and predict the impact of direct or alternating stimulation currents (intensity, frequency, duration, repetition) and resulting fields (distribution and orientation w.r.t. neurons) onto brain systems. In this section, we review some recent attempts (5 past years) to model the coupling between exogenous E-fields and neuron membranes as well as attempts to describe subsequent short-term (polarization-related) and long-term (plasticity-related) effects.

Regarding acute effects at single cell level, the classical theoretical framework to describe transmembrane potential changes generated by exogenous E-fields is provided by the cable theory (Rall 1959) also referred to as the Goldman–Hodgkin–Huxley–Katz model (GHHK) (Fig 1D). Recently, Wang et al. (Wang, Aberra et al. 2018) revisited the cable equation (CE) in order to better capture the mutual interactions between the applied field and neuron membrane polarization. Starting from previous work (Krassowska and Neu 1994), they derived new equations describing the two distinct time scales of cell membrane changes in response to an imposed extracellular E-field. Specifically, their model now accounts for initial polarization, a fast response due to transverse polarization (TP) in addition to the slower change in the mean membrane potential accounted for by the classical CE. Interestingly, authors emphasize the potential benefit of the modified CE i) in situations involving low spatial gradient of E-fields and ii) for modeling of the network behavior in response to TP-induced subthreshold modulation. A second interesting modeling study is that reported by Lafon et al. (Lafon, Rahman et al. 2017) who made use of a computational neuron model to analyze the alteration of the input(synaptic)–output(spiking) function achieved by direct current stimulation (DCS). Starting from a reduced two-compartment pyramidal cell model based on Hodgkin–Huxley formalism for active conductances, results show that a leftward shift in the input–output threshold under anodal stimulation which leads to membrane hyperpolarization at the dendrites concomitantly with depolarization at the soma. Results also suggest that synergetic interactions between the two compartments is crucial for describing the effects of DCS at cellular level. Finally, it is worth mentioning that the GHHK theory assumes ionic homogeneity and thus breaks down for small neuronal compartments such as dendritic spines or synaptic terminals that are likely impacted by weak E-Fields. Readers may refer to a broader theory based on Poisson-Nernst-Planck (PNP) approximation to more accurately account for electrodiffusion and in neuronal nanostructures (review in (Holcman and Yuste 2015)).

So far, only few studies have addressed the modeling of acute effects of low-magnitude E-fields at the neuronal population level. One study attempted, using the neural mass formalism, to describe the impact of the electric field induced by tDCS on evoked brain responses (Molaee-Ardekani, Marquez-Ruiz et al. 2013). Using the “lambda-E” model {Ruffini, 2013 #44} inspired by earlier in vitro work, a model featuring pyramidal cells, slow and fast GABAergic interneurons, was used to fit airpuff evoked responses in the rabbit somatosensory cortex (Marquez-Ruiz, Leal-Campanario et al. 2012). Different physiological mechanisms were tested, and the optimal fit between simulated and experimental evoked responses was obtained for a depolarization of pyramidal cells and slowkinetics interneurons (SOM+) combined with a hyperpolarization of fast-kinetics interneurons (basket cells). Interestingly, this modeling study emphasizes that low-magnitude electric fields as induced by tDCS interact with all types of neurons and have differential effects depending on their typology and orientation in the cortex. Future computational studies should also explore possible synapseindependent ephaptic effects (Anastassiou, Perin et al. 2011) according to which “tDCS-activated” neurons generate electric field interactions with neighbor (i.e. juxtaposed) “passive” neurons (Fig 1D). Such interactions were hypothesized to reinforce synchronization of neuronal activity and thus impact information processing and plasticity in the brain.

Finally, long-term effects induced by tDCS-level electric fields are a topic of outmost importance, especially in the context of neurological disorders like epilepsy where long-lasting “protective” (i.e. aimed at reducing cortical excitability) effects should be maximized. Among the limited number of studies reported so far, one can quote the neural mass model proposed by Modolo et al. (Modolo, Thomas et al. 2013) featuring a simplified calcium-dependent synaptic plasticity mechanism, supported by the evidence that post-synaptic calcium concentration regulates the trafficking of AMPA receptors (Shouval, Castellani et al. 2002). In this model, post-synaptic calcium concentration was expressed as a low-pass filtered version of the EEG, with a time constant derived from experimental values for calcium concentration changes(Manahan-Vaughan, Kulla et al. 2000). Testing different scenarios for the possible physiological targets of the induced electric field, it was shown that levels of membrane depolarization on the order of 0.5 mV could be sufficient to induce significant modulations of alpha power in simulated EEG signals.

5. Conclusion

The multitude of neurophysiological effects reported by the use of low-magnitude electric fields (on the order of 1 V/m) has sparked a significant interest due to its outstanding potential to offer novel non-invasive therapies for neurological disorders. However, the critical review of this literature highlights some crucial points, that should be carefully addressed in future studies tackling this question, not only to improve the interpretation of results in this research field, but also to bridge the gap between in vitro/in vivo data and reported behavioral effects. First, a systematic dosimetry of the electric field delivered within each experimental setup should be systematically provided as a good practice. Failing to provide such information might complicate the comparison between studies, and induce confusion regarding the possible targets and mechanisms. Second, the order of magnitude of the electric field used in vitro should be as close as possible to the fields induced by tDCS/tACS, in order to provide meaningful insights into the results obtained clinically. One problem is indeed that most in vitro studies investigating interaction mechanisms use electric fields that are 10 to 100 times higher than those generated by tCS. Third, combined in vivo/in silico studies should be encouraged as an attempt to validate candidate interaction mechanisms, which could then be exploit to unleash the full potential of tDCS/tACS for therapy. Finally, while neurons appear as the most likely candidate to explain the effects of low-level electric fields, it might be time to explore beyond this paradigm to consider potential effects on astrocytes or microglia. While experimentally challenging, this might prove extremely fruitful to reach a proper understanding of the involved phenomena.

Acknowledgement

Funding from: (1) European Community, Project “Luminous”, Future Emerging Technologies (H2020FETOPEN-2014–2015-RIA), agreement No. 686764 and (2) NIH (R01 grant), Application Number: R01NS092760–01A1.

Footnotes

Disclosure

The authors declare no conflict of interest.

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.

References

  1. Anastassiou CA, Perin R, Markram H and Koch C (2011). “Ephaptic coupling of cortical neurons.” Nat Neurosci 14(2): 217–223. [DOI] [PubMed] [Google Scholar]
  2. Attwell D (2003). “Interaction of low frequency electric fields with the nervous system: the retina as a model system.” Radiat Prot Dosimetry 106(4): 341–348. [DOI] [PubMed] [Google Scholar]
  3. Bikson M, Inoue M, Akiyama H, Deans JK, Fox JE, Miyakawa H and Jefferys JG (2004). “Effects of uniform extracellular DC electric fields on excitability in rat hippocampal slices in vitro.” J Physiol 557(Pt 1): 175190. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Datta A, Bansal V, Diaz J, Patel J, Reato D and Bikson M (2009). “Gyri-precise head model of transcranial direct current stimulation: improved spatial focality using a ring electrode versus conventional rectangular pad.” Brain Stimul 2(4): 201–207, 207 e201. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Debanne D, Campanac E, Bialowas A, Carlier E and Alcaraz G (2011). “Axon physiology.” Physiol Rev 91(2): 555–602. [DOI] [PubMed] [Google Scholar]
  6. Fritsch B, Reis J, Martinowich K, Schambra HM, Ji Y, Cohen LG and Lu B (2010). “Direct current stimulation promotes BDNF-dependent synaptic plasticity: potential implications for motor learning.” Neuron 66(2): 198–204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Frohlich F and McCormick DA (2010). “Endogenous electric fields may guide neocortical network activity.” Neuron 67(1): 129–143. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Gellner AK, Reis J and Fritsch B (2016). “Glia: A Neglected Player in Non-invasive Direct Current Brain Stimulation.” Front Cell Neurosci 10: 188. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Hirata A, Takano Y, Fujiwara O, Dovan T and Kavet R (2011). “An electric field induced in the retina and brain at threshold magnetic flux density causing magnetophosphenes.” Phys Med Biol 56(13): 4091–4101. [DOI] [PubMed] [Google Scholar]
  10. Holcman D and Yuste R (2015). “The new nanophysiology: regulation of ionic flow in neuronal subcompartments.” Nat Rev Neurosci 16(11): 685–692. [DOI] [PubMed] [Google Scholar]
  11. Huang R, Peng L and Hertz L (1997). “Effects of a low-voltage static electric field on energy metabolism in astrocytes.” Bioelectromagnetics 18(1): 77–80. [DOI] [PubMed] [Google Scholar]
  12. IEEE (2010). “IEEE Recommended Practice for Measurements and Computations of Electric, Magnetic, and Electromagnetic Fields with Respect to Human Exposure to Such Fields, 0 Hz to 100 kHz.” IEEE Std C95.3.1–2010. [Google Scholar]
  13. International Commission on Non-Ionizing Radiation, P. (2010). “Guidelines for limiting exposure to time-varying electric and magnetic fields (1 Hz to 100 kHz).” Health Phys 99(6): 818–836. [DOI] [PubMed] [Google Scholar]
  14. Jackson MP, Rahman A, Lafon B, Kronberg G, Ling D, Parra LC and Bikson M (2016). “Animal models of transcranial direct current stimulation: Methods and mechanisms.” Clin Neurophysiol 127(11): 34253454. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Jefferys JG, Deans J, Bikson M and Fox J (2003). “Effects of weak electric fields on the activity of neurons and neuronal networks.” Radiat Prot Dosimetry 106(4): 321–323. [DOI] [PubMed] [Google Scholar]
  16. Kabakov AY, Muller PA, Pascual-Leone A, Jensen FE and Rotenberg A (2012). “Contribution of axonal orientation to pathway-dependent modulation of excitatory transmission by direct current stimulation in isolated rat hippocampus.” J Neurophysiol 107(7): 1881–1889. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Krassowska W and Neu JC (1994). “Response of a single cell to an external electric field.” Biophys J 66(6): 1768–1776. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Kronberg G, Bridi M, Abel T, Bikson M and Parra LC (2017). “Direct Current Stimulation Modulates LTP and LTD: Activity Dependence and Dendritic Effects.” Brain Stimul 10(1): 51–58. •• Experimental demonstration that DC currents can modulate synaptic plasticity. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Laakso I and Hirata A (2012). “Computational analysis of thresholds for magnetophosphenes.” Phys Med Biol 57(19): 6147–6165. [DOI] [PubMed] [Google Scholar]
  20. Lafon B, Rahman A, Bikson M and Parra LC (2017). “Direct Current Stimulation Alters Neuronal Input/Output Function.” Brain Stimul 10(1): 36–45. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Lovsund P, Oberg PA, Nilsson SE and Reuter T (1980). “Magnetophosphenes: a quantitative analysis of thresholds.” Med Biol Eng Comput 18(3): 326–334. [DOI] [PubMed] [Google Scholar]
  22. Manahan-Vaughan D, Kulla A and Frey JU (2000). “Requirement of translation but not transcription for the maintenance of long-term depression in the CA1 region of freely moving rats.” J Neurosci 20(22): 85728576. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Marquez-Ruiz J, Leal-Campanario R, Sanchez-Campusano R, Molaee-Ardekani B, Wendling F, Miranda PC, Ruffini G, Gruart A and Delgado-Garcia JM (2012). “Transcranial direct-current stimulation modulates synaptic mechanisms involved in associative learning in behaving rabbits.” Proc Natl Acad Sci U S A 109(17): 6710–6715. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Mina F, Modolo J, Recher F, Dieuset G, Biraben A, Benquet P and Wendling F (2017). “Model-guided control of hippocampal discharges by local direct current stimulation.” Sci Rep 7(1): 1708. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Miranda PC, Lomarev M and Hallett M (2006). “Modeling the current distribution during transcranial direct current stimulation.” Clin Neurophysiol 117(7): 1623–1629. [DOI] [PubMed] [Google Scholar]
  26. Modolo J, Thomas AW and Legros A (2013). “Neural mass modeling of power-line magnetic fields effects on brain activity.” Front Comput Neurosci 7: 34. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Molaee-Ardekani B, Marquez-Ruiz J, Merlet I, Leal-Campanario R, Gruart A, Sanchez-Campusano R, Birot G, Ruffini G, Delgado-Garcia JM and Wendling F (2013). “Effects of transcranial Direct Current Stimulation (tDCS) on cortical activity: a computational modeling study.” Brain Stimul 6(1): 25–39. •• Computational modeling study providing evidence that externally-applied fields also affect the activity of interneurons. [DOI] [PubMed] [Google Scholar]
  28. Monai H and Hirase H (2018). “Astrocytes as a target of transcranial direct current stimulation (tDCS) to treat depression.” Neurosci Res 126: 15–21. [DOI] [PubMed] [Google Scholar]
  29. Monai H, Ohkura M, Tanaka M, Oe Y, Konno A, Hirai H, Mikoshiba K, Itohara S, Nakai J, Iwai Y and Hirase H (2016). “Calcium imaging reveals glial involvement in transcranial direct current stimulationinduced plasticity in mouse brain.” Nat Commun 7: 11100 •• Evidence from in vivo recordings that glia is also involved in the physiological mechanisms of tCS. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Podda MV, Cocco S, Mastrodonato A, Fusco S, Leone L, Barbati SA, Colussi C, Ripoli C and Grassi C (2016). “Anodal transcranial direct current stimulation boosts synaptic plasticity and memory in mice via epigenetic regulation of Bdnf expression.” Sci Rep 6: 22180. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Rahman A, Lafon B, Parra LC and Bikson M (2017). “Direct current stimulation boosts synaptic gain and cooperativity in vitro.” J Physiol 595(11): 3535–3547. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Rahman A, Reato D, Arlotti M, Gasca F, Datta A, Parra LC and Bikson M (2013). “Cellular effects of acute direct current stimulation: somatic and synaptic terminal effects.” J Physiol 591(10): 2563–2578. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Rall W (1959). “Branching dendritic trees and motoneuron membrane resistivity.” Exp Neurol 1: 491–527. [DOI] [PubMed] [Google Scholar]
  34. Reato D, Rahman A, Bikson M and Parra LC (2010). “Low-intensity electrical stimulation affects network dynamics by modulating population rate and spike timing.” J Neurosci 30(45): 15067–15079. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Rush S and Driscoll AM (1968). “Current Distribution in the Brain From Surface Electrodes.” Anesthesia & Analgesia 47(6): 717–723. [PubMed] [Google Scholar]
  36. Sadleir RJ, Vannorsdall TD, Schretlen DJ and Gordon B (2010). “Transcranial direct current stimulation (tDCS) in a realistic head model.” Neuroimage 51(4): 1310–1318. [DOI] [PubMed] [Google Scholar]
  37. Shouval HZ, Castellani GC, Blais BS, Yeung LC and Cooper LN (2002). “Converging evidence for a simplified biophysical model of synaptic plasticity.” Biol Cybern 87(5–6): 383–391. [DOI] [PubMed] [Google Scholar]
  38. Stafford J, Brownlow ML, Qualley A and Jankord R (2018). “AMPA receptor translocation and phosphorylation are induced by transcranial direct current stimulation in rats.” Neurobiol Learn Mem 150: 36–41. [DOI] [PubMed] [Google Scholar]
  39. Sun Y, Lipton JO, Boyle LM, Madsen JR, Goldenberg MC, Pascual-Leone A, Sahin M and Rotenberg A (2016). “Direct current stimulation induces mGluR5-dependent neocortical plasticity.” Ann Neurol 80(2): 233–246. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Voroslakos M, Takeuchi Y, Brinyiczki K, Zombori T, Oliva A, Fernandez-Ruiz A, Kozak G, Kincses ZT, Ivanyi B, Buzsaki G and Berenyi A (2018). “Direct effects of transcranial electric stimulation on brain circuits in rats and humans.” Nat Commun 9(1): 483 •• Intracranial recordings of E-fields induced by tCS in human cadavers with a precise evaluation of current shunting by the skin, skull and CSF. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Wang B, Aberra AS, Grill WM and Peterchev AV (2018). “Modified cable equation incorporating transverse polarization of neuronal membranes for accurate coupling of electric fields.” J Neural Eng 15(2): 026003. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES