Sunteți pe pagina 1din 14

NeuroImage 87 (2014) 332–344

Contents lists available at ScienceDirect

NeuroImage
journal homepage: www.elsevier.com/locate/ynimg

A computational modelling study of transcranial direct current


stimulation montages used in depression
Siwei Bai a, Socrates Dokos a, Kerrie-Anne Ho b,d, Colleen Loo b,c,d,⁎
a
Graduate School of Biomedical Engineering, Faculty of Engineering, University of New South Wales (UNSW), NSW 2052, Australia
b
School of Psychiatry, UNSW, NSW 2052, Australia
c
Department of Psychiatry, St George Hospital, NSW 2217, Australia
d
Black Dog Institute, NSW 2031, Australia

a r t i c l e i n f o a b s t r a c t

Article history: Transcranial direct current stimulation (tDCS) is a neuromodulatory technique which involves passing a mild
Accepted 5 November 2013 electric current to the brain through electrodes placed on the scalp. Several clinical studies suggest that tDCS
Available online 15 November 2013 may have clinically meaningful efficacy in the treatment of depression. The objective of this study was to simulate
and compare the effects of several tDCS montages either used in clinical trials or proposed, for the treatment of
Keywords:
depression, in different high-resolution anatomically-accurate head models. Detailed segmented finite element
Transcranial direct current stimulation
Depression
head models of two subjects were presented, and a total of eleven tDCS electrode montages were simulated.
Computational model Sensitivity analysis on the effects of changing the size of the anode, rotating both electrodes and displacing the
anode was also conducted on selected montages. The F3–F8 and F3–F4 montages have been used in clinical trials
reporting significant antidepressant effects and both result in relatively high electric fields in dorsolateral
prefrontal cortices. Other montages using a fronto-extracephalic or fronto-occipital approach result in greater
stimulation of central structures (e.g. anterior cingulate cortex) which may be advantageous in treating depres-
sion, but their efficacy has yet to be tested in randomised controlled trials. Results from sensitivity analysis
suggest that electrode position and size may be adjusted slightly to accommodate other priorities, such as skin
discomfort and damage.
© 2013 Elsevier Inc. All rights reserved.

Introduction meaningful efficacy (Boggio et al., 2008; Brunoni et al., 2011; Fregni
et al., 2006a, 2006b; Kalu et al., 2012; Loo et al., 2012; Martin et al.,
Transcranial direct current stimulation (tDCS) is a neuromodulatory 2011; Palm et al., 2011). These studies focused on anodal stimulation
technique which involves passing a mild electric current to the brain of the left dorsolateral prefrontal cortex (DLPFC), based on observations
through electrodes placed on the scalp. This direct constant flow of that this area has been associated with underactivity in depression
current modulates underlying cortical activity with specific outcomes (Grimm et al., 2008). However, studies differed in the location of the
related to anodal or cathodal stimulation (Nitsche and Paulus, 2000, cathode, i.e. the return electrode — right supraorbital, right lateral
2001). The relative position (electrode montage) and size of the anode orbitofrontal, right DLPFC or in an extracephalic position. Though the
and cathode determine the distribution of current density throughout anodal left DLPFC electrode is often considered the “active” electrode,
the brain (Bikson et al., 2010; Datta et al., 2011; Lee et al., 2012; the placement of the cathode is important for several reasons: shunting
Miranda et al., 2009; Wagner et al., 2007). Thus there is potential for of much of the current over the scalp may occur if the inter-electrode
stimulation to be focussed on specific cortical brain regions for thera- distance is too close (Datta et al., 2008; Miranda et al., 2006; Weaver
peutic or investigative purposes or more diffuse effects can be produced et al., 1976), current density under the anode is affected by the place-
if widespread activation of brain regions is desired. ment of the reference or “return” electrode (Bikson et al., 2010; Datta
A key application of tDCS has been investigated in the treatment of et al., 2011), and the pattern of brain areas stimulated will be deter-
depression. Several recent open label and placebo-controlled trials, mined by the overall montage. All of these factors may have important
and a meta-analysis of mean change in depression scores from therapeutic implications.
placebo-controlled studies suggest that tDCS may have clinically Pathophysiological changes in depression are system-wide, involv-
ing a network of various cortical and limbic structures rather than a sol-
itary brain region such as the left DLPFC (Mayberg, 2007). Hypoactivity
⁎ Corresponding author at: Black Dog Institute, Hospital Road, Prince of Wales Hospital,
Randwick, NSW 2031, Australia. Fax: +61 2 9113 3734.
in cortical regions and hyperactivity in subcortical and limbic regions is
E-mail addresses: s.bai@unsw.edu.au (S. Bai), s.dokos@unsw.edu.au (S. Dokos), often associated with symptoms of depression (Fitzgerald et al., 2008;
kerrie-anne.ho@unsw.edu.au (K.-A. Ho), colleen.loo@unsw.edu.au (C. Loo). Mayberg, 1997). Meta-analyses have identified frontal and temporal

1053-8119/$ – see front matter © 2013 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.neuroimage.2013.11.015
S. Bai et al. / NeuroImage 87 (2014) 332–344 333

cortices, the insula and cerebellum as regions of hypoactivity while Asian male whose MRI head scan, labelled “Msub” (short for male
subcortical and limbic regions tend to be hyperactive. This distributed subject), was truncated at the level of cervical vertebra 6. The other
network of structures includes the DLPFC, medial prefrontal cortex was a 42-year-old Caucasian female, labelled “Fsub” (female subject,
(MPFC), orbitofrontal cortex (OFC), as well as the anterior cingulate Fig. 1S in Supplementary data): her scan was truncated at the level
cortex (ACC), insula and hippocampus (Fox et al., 2012; Mayberg, of the atlas-axis, i.e., cervical vertebrae 1–2. T1-weighted MRI
2003). Most recently, functional connectivity studies have suggested scans of both subjects were obtained from Neuroscience Research
altered activity at a network level during the resting state (Carballedo Australia. The scans were sagittally-oriented with voxel resolution
et al., 2011). In particular, there is increased functional connectivity in of 1 mm × 1 mm × 1 mm. The images of Msub were later down-
the subgenual anterior cingulate (sgACC), thalamus and OFC in people sampled to 1.5 mm in every dimension.
with depression (Greicius et al., 2007). Further, overactivity in the Head tissue masks were obtained using a combination of automated
sgACC has been shown to be strongly negatively correlated with resting and manual segmentation softwares. Automated mask generation was
state underactivity in the left DLPFC (Fox et al., 2012). performed using BrainSuite, an open-source package from the Laborato-
Studies of deep brain stimulation (DBS) in depression have also ry of NeuroImaging at the University of California (Shattuck and Leahy,
provided insight into the critical regions involved in depression. Consis- 2002). Thus, tissue compartments including skin, skull, cerebrospinal
tent with imaging studies, DBS interventions targeted at the sgACC have fluid (CSF), grey matter (GM) and white matter (WM) were generated
demonstrated efficacy in reducing symptoms of depression (Lozano from the MRI data. The segmented masks were exported from
et al., 2012; Mayberg et al., 2005). DBS to specific regions of the basal BrainSuite as grayscale images, and imported into ScanIP (Simpleware
ganglia such as the nucleus accumbens (NAcc) and the ventral Ltd., UK) for manual correction and further processing. The five original
capsule/ventral striatum (VC/VS) have also been found to have signifi- masks were hence divided into more compartments:
cant antidepressant effects (Anderson et al., 2012; Bewernick et al.,
• masks representing eyes, paranasal sinuses, larynx and cervical verte-
2010, 2012; Malone et al., 2009).
brae were separated from the skin and skull, as shown in Fig. 1a. In
As the therapeutic potential of tDCS in psychiatric disorders is
addition, the major foramina of the skull were included in the skull
further explored, information on how different electrode arrangements
mask, including the superior orbital fissure, optic canal, foramen
determine current density in key brain regions, is essential. This study
ovale and foramen magnum;
compared the effects of several DCS montages, with realistic head
• the skull was divided into the cranium and jaw. The cranium was then
models reconstructed from MRI head scans, by investigating the brain
subdivided into three layers, with spongy bone tissue as the middle
electric field (E-field) distribution and the average E-field in various
layer, and compact bone tissue as the outermost and innermost layers.
brain regions. tDCS montages modelled were those used in recent
The jaw was considered compact. These skull compartments are
tDCS depression studies: the F3–supraorbital (F3–SO) montage first
shown in Fig. 1b;
used when interest was rekindled in tDCS from 2006 onwards (Boggio
• the brain masks consisted of GM, WM, cerebellum (CB, with
et al., 2008; Fregni et al., 2006a, 2006b; Loo et al., 2010; Palm et al.,
brainstem) and the cervical spinal cord (SC), as well as the ventricular
2011), and modified approaches in which the cathode was moved
system which was later assigned to the CSF mask;
more laterally to reduce shunting, F3–F8 (Loo et al., 2012), to the right
• several brain regions of interest (ROIs), considered important in tDCS
DLPFC, F3–F4 (Brunoni et al., 2011, 2013; Dell'Osso et al., 2012;
therapeutic effects, were further segmented from the GM mask as
Ferrucci et al., 2009a, 2009b), or to an extracephalic position to achieve
shown in Fig. 1c — anterior cingulate cortices (ACCs), amygdalae, hip-
a more widespread pattern of brain activation, F3–extracephalic
pocampi, dorsolateral prefrontal cortices (DLPFCs) and orbitofrontal
(F3–EC, brain sites based on the 10–20 EEG system; Martin et al.,
cortices (OFCs).
2011). The bilateral supraorbital–extracephalic (SO–EC) montage
most commonly used in earlier, pre-2000 studies, involving two In the head models, fat and muscle were included in the skin com-
small anodes at the frontal poles and an extracephalic cathode was partment, due to the fact that their conductivities are of the same
also modelled (Arul-Anandam and Loo, 2009; Lippold and Redfearn, order of magnitude as skin conductivity (Gabriel et al., 1996). Similarly,
1964; Redfearn et al., 1964). In addition, several hypothetical the venous sinuses and cranial arteries were included in the cerebrospi-
montages were modelled: supraorbital–occipital (SO–OCC), pre- nal fluid compartment. Finally, any remaining blank voxels were manu-
mised on maximal stimulation of the sgACC and other central and ally assigned to the most appropriate neighbouring mask.
midline subcortical structures; temporal–extracephalic (TMP–EC), To examine the effect of an extracephalic clinical electrode montage
prioritising temporal lobe stimulation as neurotrophic changes in used in some tDCS studies (Martin et al., 2011; Moliadze et al., 2010), a
this region may have a key role in the pathophysiology of depression synthetic upper torso attached to the segmented head was manually
(Pittenger and Duman, 2007), and supraorbital–cerebellum (SO– painted in ScanIP up to the level above the axilla based on anthropomet-
CB), as abnormal cerebellar modulation of the cerebello-thalamo- ric measurements (Dreyfuss and Tilley, 1993), as shown in Fig. 1d. A
cortical pathway has been implicated in the mood and cognitive similar approach was used in other studies (Borckardt et al., 2012;
symptoms associated with several psychiatric disorders, including Datta et al., 2011, 2012; Mendonca et al., 2011).
bipolar disorder and depression (Hoppenbrouwers et al., 2008). The + FE Free meshing algorithm in the + FE module of ScanIP
The montages were modelled in two subjects — one male and one (v4.3) was selected to generate the tetrahedral mesh elements for the
female — to examine the extent to which inter-individual differences high-resolution head models, with a compound coarseness of -30.
in head anatomy affect variation in electric field with different The meshes were then imported into the COMSOL Multiphysics FE
montages. Finally, a sensitivity analysis was performed to examine the solver (v4.2).
effects of displacing the anodal electrode by ~ 1 cm, to inform on
the likely importance of accuracy in electrode placement in clinical Tissue conductivities
applications.
Most compartments of the head models were considered to be elec-
Methods trically homogeneous and isotropic. The electrical conductivity of
paranasal sinuses (and larynx) was set to zero. Conductivities of the
Image segmentation and mesh generation scalp, compact and spongy bones of the skull, CSF, GM and WM, were
assigned mean values from multiple studies (Akhtari et al., 2000,
Two different high-resolution computational head models were 2002; Baumann et al., 1997; Geddes and Baker, 1967; Gonçalves et al.,
reconstructed from human subjects. One subject was a 35-year-old 2003; Gutierrez et al., 2004; Lai et al., 2005; Oostendorp et al., 2000).
334 S. Bai et al. / NeuroImage 87 (2014) 332–344

brain

cerebrospinal
fluid
a)
skull
eye
sinuses
& pharynx

skin vertebrae

spongy bone tissue GM


WM
ACC

DLPFC
b) c) OFC

AH CB
compact bone tissue
& the jaw SC

d)

Fig. 1. a): Segmentation of the head model “Msub”: skin, eyes, paranasal sinuses (with larynx), skull (including compact bone tissue and spongy bone tissue), vertebrae, CSF and brain.
b): Segmentation of skull: compact bone tissue and spongy bone tissue. c): Detailed segmentation of the brain, including defined regions for white matter (WM), grey matter (GM),
anterior cingulate cortex, (ACC), dorsolateral prefrontal cortex (DLPFC), orbitofrontal cortex (OFC), amygdala and hippocampus (AH), cerebellum (CB, with brainstem), and cervical spinal
cord (SC). d): Frontal view of the model with extended shoulder. To respect the subject's privacy, the eyes of the model are hidden.

All conductivity values are listed in Table 1. The conductivities of the


eyes and the synthetic torso were assigned to the scalp conductivity.

White matter conductivity anisotropy Table 1


Tissue conductivities.
Modelling studies dedicated to the comparison between isotropic Compartment Electrical conductivity (S/m)
and anisotropic conductivities (Lee et al., 2012; Shahid et al., 2013;
Scalp 0.41
Suh et al., 2012), have shown that the presence of WM anisotropy
Eyes 0.41
resulted in a significant difference in regional E-fields, especially in the Sinus 0
deep brain structures. Hence, the WM anisotropic conductivity was CSF 1.79
also adopted in this study. GM 0.31
WM 0.14
DT-MRI was performed only on Msub in 61 gradient directions, with
WM (longitudinal) 0.65
voxel resolution of 2.5 mm × 2.5 mm × 2.5 mm. After being registered WM (transverse) 0.065
to the T1 structural scan in Amira (Visage Imaging GmbH, Germany), Vertebrae 0.013
diffusion tensor calculation was performed in FSL, an open source soft- Synthetic torso 0.41
ware developed by the FMRIB Analysis Group of University of Oxford Skull (compact bone) 0.006
Skull (spongy bone) 0.028
(Behrens et al., 2003a, 2003b, 2007). Eigenvectors and fractional
S. Bai et al. / NeuroImage 87 (2014) 332–344 335

anisotropy (FA) were then calculated, with the latter being widely used Anode Cathode
to denote the degree of anisotropy: typically greater than 0.45 for WM
(Johansen-Berg and Behrens, 2009).
The conductivity tensor of WM, σ, was calculated from:

σ ¼ Sdiagðσ l ; σ t ; σ t ÞS ; ð1Þ
F3-SO

where S is the orthogonal matrix of unit eigenvectors obtained from the


WM diffusion tensor, and σl and σt are the assigned conductivities lon-
gitudinal and transverse to the fibre directions respectively, with
σl : σt = 10 : 1 (Nicholson, 1965). σl and σt were calculated using the
volume-constraint method (Wolters et al., 2006). This resulted in the
values of WM longitudinal and transverse conductivities being 0.65
and 0.065 S/m respectively. F3-F8
Following the diffusion tensor calculation, the calculated conduc-
tivity tensors of data points in the DT-MRI scans were then linked
to their individual coordinates in the Msub model. This process
was performed using MATLAB software (The Mathworks, USA).
Only fibre conductivity data having a strong anisotropy signal
(FA ≥ 0.45) were exported.

Boundary conditions for volume conductor model sF3-F8


All head compartments in the tDCS simulations were formulated as
passive volume conductors using Laplace's equation:

∇  ð−σ∇φÞ ¼ 0; ð2Þ

where φ is the electric potential, and σ is the conductivity tensor.


For scalp boundaries at the electrodes, two types of boundary condi-
tions were modelled separately: F3-F4-1
• normal component of inward current density set to Jn for anode, and
− Jn for cathode, where J n ¼ area of I electrode, with Is defined as the applied
s

stimulus current fixed at a DC level of 1 mA;


• constant voltage set to V for anode, and − V for cathode, with V satis-
fying ∫ J n dS ¼ I s , where S was the electrode area.
S
These two types of boundary conditions represented two extreme
cases for the electrode. In clinical reality, typical use of a saline-soaked F3-F4-2
sponge between the skin and electrode pads suggests that neither
constant current, or constant voltage conditions are likely to precisely
hold at the scalp–electrode interface. Nonetheless, our simulation
results suggested that E-field distributions remained generally the
same regardless of the choice of electrode boundary condition, even
though absolute E-field values were slightly different between the two Fig. 2. tDCS electrode placement — part 1: F3-supraorbital (F3–SO), F3–F8, sF3–F8, F3–F4-
1 and F3–F4-2. The red and blue electrodes represent the anode and cathode, respectively.
cases. As a result, we report only results adopting constant current To respect the subject's privacy, the eyes of the model are hidden.
electrode boundaries. Results using fixed voltage electrode boundaries
are supplied in the Supplementary data (Fig. 2S).
The rest of the boundary conditions were:
• ground (zero electric potential) at the lower boundary of the synthetic the occiput for the SO–OCC montage, as the aim was to tailor current
(extended) torso; pathways to broadly target the medial subcortical structures described
• all other external boundaries were assigned as electric insulators above (sgACC, basal ganglia) rather than deliver occipital stimula-
(zero normal component of current density); tion. The electrode placements for these montages are described as
• continuous current density (i.e., flux continuity) across all interior follows:
boundaries.
• F3–SO: both electrodes were 7 cm × 5 cm rectangular pads. The
Electrode placements anode was placed over the F3 electrode site on a standard 10–20
EEG cap system, with the long axis of the pad pointing towards
Electrodes located on the scalp surface were defined mathematical- the vertex. The cathode was placed above the arcus superciliaris
ly, which enabled their size, location and orientation to be readily on the right, with the long axis of the pad parallel to the horizontal
adjusted. A total of eleven tDCS electrode montages were simulated, plane.
as shown in Figs. 2 and 3. For most montages, standard 7 cm × 5 cm • F3–F8: both electrodes were 7 cm × 5 cm rectangular pads. The
electrodes were used. For montages with the return electrode placed anode was placed at the same location as in F3–SO. The cathode
in an EC position, focal stimulation of the extracephalic region was not was placed over the F8 EEG cap electrode site, with the lower
required and a 10 cm × 10 cm electrode was used for reasons of edge of the pad 4–5 mm below the eye socket, and the long axis
comfort. Similarly, a 10 cm × 10 cm return electrode was placed over perpendicular to the horizontal plane.
336 S. Bai et al. / NeuroImage 87 (2014) 332–344

Anode Cathode • F3–F4-2: both electrodes were 7 cm × 5 cm rectangular pads, and


were placed at the same sites as the F3–F4-1 configuration, but the
orientations of both electrodes were perpendicular to those in F3–
F4-1.
• F3–EC: the 7 cm × 5 cm anode was placed at the same location as
F3-EC
in F3–SO, and the cathode was a 10 cm × 10 cm rectangular pad
placed on the right shoulder.
• sF3–EC: the anode was placed at the same location as that of F3–EC,
but the size of the pad was 4 cm × 4 cm. The 10 cm × 10 cm cath-
ode was placed in the same location as the F3–EC configuration.
• TMP–EC: the anode was a 7 cm × 5 cm rectangular pad placed
between the F3 and T3 EEG positions, with the long axis of the
pad parallel to the line connecting F3 and T3. The 10 cm × 10 cm
sF3-EC cathode was placed on the right shoulder.
• SO–OCC: the anode was a 7 cm × 5 cm rectangular pad, placed
above the arcus superciliaris on the left, with the long axis of
the pad parallel to the horizontal plane. The cathode was a
10 cm × 10 cm square pad centred over the inion, with the larger
electrode size used to diffuse cathodal effects.
• SO–CB: the anode was a 7 cm × 5 cm rectangular pad, placed at
the same position as the SO–OCC configuration. The cathode was
TMP-EC a 10 cm × 5 cm rectangular pad, with the top edge centred at the
inion, and the long axis parallel to the horizontal plane, to achieve
bilateral cerebellar stimulation.
• SO–EC: two circular anodes with a diameter of 0.5 in. (1.27 cm)
were placed above the arcus superciliaris on both sides, with the
cathode located as in the F3–EC configuration.

Data analysis
SO-OCC
Three types of head models were investigated in this study —
“Msub-aniso” (referring to the Msub model with WM anisotropy),
“Msub” (referring to the Msub model without WM anisotropy) and
“Fsub” (without WM anisotropy). The models were solved using the
segregated numerical solver in COMSOL (v4.2) on a Windows 64-bit
workstation with 24 GB RAM utilising 4 processors. To solve the station-
ary equations, a direct linear solver was utilised with an absolute error
tolerance set to 10−5. It took ~20 min to solve for each simulation, for
SO-CB approximately 1.4 × 106 degrees of freedom.
Simulation results were analysed by comparing the difference in
brain E-field distribution among the various electrode montages and
different head models. The analysis also focused on comparing the
average E-field magnitude E in several ROIs in the brain. E was calculated
using:

Z
SO-EC jEj dV

E¼ Z V
; ð3Þ
dV
V

Fig. 3. tDCS electrode placement — part 2: F3-extracephalic (F3–EC), sF3–EC, temporal– where |E| is the E-field magnitude in the ROI in question, and ∫
extracephalic (TMP–EC), supraorbital–occipital (SO–OCC), supraorbital–cerebellum (SO– is a volume integral over this region. Note that the denominator V
CB) and supraorbital–extracephalic (SO–EC). The red and blue electrodes represent the
anode and cathode, respectively. To respect the subject's privacy, the eyes of the model
is simply the volume of the ROI. The mean of E and its standard
are hidden. error (SE) were then determined across all three head models.
The difference in the E-field was considered significant, if the
E-field range (mean ± SE) between any two montages did not
overlap.
In addition, a sensitivity analysis on E-fields in the ROIs to electrode
• sF3–F8: the anode was placed at the same F3–F8 location, but the displacement from the original position was conducted on the Msub-
size of the anode pad was 4 cm × 4 cm. The cathode was the aniso. Three montages were included in this analysis: F3–F4-1, F3–F8
same as in F3–F8. and F3–EC. For each montage, the anode F3 was perturbed from its
• F3–F4-1: both electrodes were 7 cm × 5 cm rectangular pads. The original position anteriorly, posteriorly, laterally or medially (note: the
anode was placed at the same location as in F3–SO, and the cathode antero-posterior displacement was on a sagittal plane, whereas the
was placed over the F4 electrode site on the 10–20 EEG cap system, medio-lateral displacement was on a horizontal plane). The displace-
with the long axis pointing towards the vertex of the head. ment was set to 1 cm, which is the likely margin of error in clinical trials.
S. Bai et al. / NeuroImage 87 (2014) 332–344 337

The absolute differences in average E-fields from the original place- distributed current density was present in the whole brain, especially
ments were calculated. in the ventral part, owing to current predominantly flowing inferiorly
and posteriorly. Compared to the other F3 montages, the E-field magni-
Results tude was weaker in the frontal lobe with F3–EC. SO–OCC and TMP–EC
had E-field characteristics similar to F3–EC, except that the predomi-
Fig. 4 shows the E-field magnitude and direction on the cortical nant E-field direction with SO–OCC was anterior–posterior, and the
surface of the brain for six selected tDCS electrode montages with E-field magnitude was large in the left temporal lobe with TMP–EC.
Msub-aniso. Fig. 5 shows the E-field profile with the same head model The mean and standard error of spatially-averaged E-field magni-
in cross-sectional slices of the brain for the selected tDCS electrode tude E for the three head models were shown in Fig. 6, with separate
montages. The montages which utilised the F3 anode and a contralateral data of E in individual head models shown in the Supplementary data
frontal cephalic cathode (F3–SO, F3–F8, F3–F4-1 as shown in Fig. 4) (Figs. 3S–5S). In comparison to the F3–SO montage, the modified mon-
exhibited higher current density predominately in the frontal lobes of tages with the anode on F3 but cathode placed at a greater distance from
the brain. F3–F8 and F3–F4-1 shared a similar E-field magnitude profile, the anode (F3–F8, F3–F4, F3–EC) all resulted in less shunting, i.e. more
but the E-field directions were quite distinctive, due to the different current entered the brain, with F3–SO significantly different from F3–
positions of the cathodes. As for F3–EC, a more uniform and widely F8 and F3–EC. The F3–EC montage resulted in the least shunting, with

F3-SO

F3-F8
V/m
≥ 0.25

F3-F4-1

0.15

F3-EC

0.05

SO-OCC

TMP-EC

Fig. 4. E-field magnitude distribution and direction in the whole brain for Msub-aniso model for six selected tDCS montages. The leftmost and two middle columns feature the lateral view
from the left, the frontal view and the top view, respectively. The black box specifies the location of the enlarged region shown in the rightmost column. The black arrows represent the
direction of E-field vectors, whereas the red arrows indicate the global electrode vector direction — from the anode to the cathode.
338 S. Bai et al. / NeuroImage 87 (2014) 332–344

F3-SO

F3-F8 V/m
≥ 0.25

F3-F4-1
0.15

F3-EC
0.05

SO-OCC

TMP-EC

Fig. 5. E-field brain magnitude distribution in Msub-aniso in cross-sections through the DLPFC (coronal), ACC (sagittal) and OFC (horizontal) with these structures outlined. Dashed lines
indicate locations of slice planes.

more current reaching brain areas compared to the other F3 montages. electrode size at the F3 site (sF3–F8, sF3–EC) and rotation of electrodes
For F3–EC, the E-field in the DLPFCs was reduced compared to other F3 by 90° (F3–F4-2) only made a small difference to modulation patterns.
montages, but this montage more strongly modulated the ACCs and The SO–OCC montage was similar to the F3–EC montage in less
cerebellum/brainstem than the other F3 montages. Variation of shunting and a higher E-field in the ACCs than other frontal montages,
S. Bai et al. / NeuroImage 87 (2014) 332–344 339

V/m F3-SO
F3-F8
0.1 sF3-F8
F3-F4-1
F3-F4-2
F3-EC
sF3-EC
TMP-EC
0 SO-OCC
Whole brain L hemisphere R hemisphere
SO-CB
SO-EC

0.1

0
CB L DLPFC R DLPFC

0.1

0
L OFC R OFC L ACC

0.1

0
R ACC L hippocampus R hippocampus

Fig. 6. Mean and standard error of spatially-averaged E-field magnitude E in various brain regions of interest across the three head models — Msub-aniso, Msub and Fsub.

though the E-field in the DLPFC was reduced. When a smaller cathode these were in slightly different locations in each head, due to the varia-
was used and centred over the cerebellum (SO–CB), the main effect tion in cortical foldings.
compared to SO–OCC was an E-field increase in more ventral structures Fig. 8 demonstrates the sensitivity analysis of E-fields (in Msub-aniso)
(hippocampus, OFC and CB), but not to a significant level. Compared with respect to displacements of the anode in the anterior, posterior,
with more recent frontal montages, the (bilateral) SO–EC resulted in lateral and medial directions. Moving the electrode by 1 cm did
less left DLPFC modulation, but greater OFC and ACC modulation not make a substantive difference. Reduction in the inter-electrode
particularly in the left hemisphere. The TMP–EC montage most strongly distance, such as medial displacement with F3–F4-1 and F3–F8, tended
modulated the marked E-field reduction in frontal structures with to result in more shunting and less current entering the brain. More
TMP–EC. significant changes (i.e., close to 10% of the original), were (ROIs): left
The pattern revealed by the average E-field (E) magnitude with OFC in F3–F4-1 with antero-posterior displacement, left and right
Msub mostly agreed with the anisotropic model. Similarly, the ROI ACCs in F3–F8 with antero-posterior displacement, and left and right
E-field data from Fsub was largely consistent with that from both ACCs in F3–EC with antero-posterior displacement.
Msub and Msub-aniso, albeit with minor variations. Namely, Fsub
indicated more significant current-shunting in montages with closer Discussion
inter-electrode scalp distances (E of whole brain, left hemisphere and CB
with F3–SO and F3–F4 in Figs. 3S–5S), smaller inter-montage differ- Clinical implications of electrode montages
ences in E for left DLPFC, as well as increased E magnitude for right
DLPFC with both F3–F8 montages. These discrepancies are likely due A number of recent tDCS studies have reported positive antidepres-
to the variation in head geometry, especially in skull size and thickness. sant effects in depressed patients (Boggio et al., 2008; Brunoni et al.,
Regardless, the ROI E-field magnitude distribution with Fsub was quite 2011, 2013; Fregni et al., 2006a, 2006b; Kalu et al., 2012; Loo et al.,
similar to that of the corresponding Msub regions (F3–F4-1 in Fig. 7), 2012; Martin et al., 2011; Palm et al., 2011). The rationale guiding the
and there were local hot-spots evident on the cortical surface, but choice of electrode montage in these trials is up-regulation of the left
340 S. Bai et al. / NeuroImage 87 (2014) 332–344

Msub Fsub

V/m
≥ 0.25

0.15

0.05

Fig. 7. Brain E-field magnitude distribution for the Msub and Fsub models using the F3–F4-1 montage. The rectangles specify the location of the enlarged cortical regions in the bottom row.

DLPFC, with the anode placed on F3 (Boggio et al., 2008; Brunoni et al., trial. It is useful to note that use of a smaller electrode at F3 increased
2011, 2013; Dell'Osso et al., 2012; Ferrucci et al., 2009a, 2009b; Fregni the E-field at the left DLPFC. This may be a useful strategy to compensate
et al., 2006a, 2006b; Loo et al., 2010, 2012; Martin et al., 2011; Palm for loss of intensity associated with moving the cathode further away.
et al., 2011). This is based on the general premise that in depressed pa- The F3–F8 and F3–F4 montages produced similar patterns of the
tients, the left DLPFC is hypoactive while the right DLPFC is hyperactive E-field in the brain, resulting in strong stimulation of the left and
(Grimm et al., 2008). The results from the present study showed that all right DLPFC and clinical trials of these montages have demonstrated
montages with the anode placed at F3 achieved this aim. antidepressant effects (Brunoni et al., 2011, 2013; Dell'Osso et al.,
Findings from the present study also indicate the importance of 2012; Ferrucci et al., 2009a; Loo et al., 2012). The role of cathodal stim-
cathodal placement as it determines the relative current distribution ulation of ventral right frontal regions, particularly achieved with the
in other brain regions, as well as the intensity of stimulation at the left F3–F8 montage, in antidepressant efficacy is unclear. The F3–F8 mon-
DLPFC. In comparison to the F3–SO montage where the electrodes are tage resulted in a significantly greater E-field in the right OFC compared
positioned close together, montages with wider spacing between the to the F3–F4 and other montages examined. A recent trial of the F3–F8
electrodes result in less shunting and more current entering the brain. montage reported clear antidepressant results, with a 50% response rate
However, E-field levels in the DLPFC also decreased as the inter- after six weeks of tDCS (Loo et al. 2012). However, the independent ef-
electrode distance increased, in line with Moliadze et al. (2010) which fects of right frontal cathodal stimulation have not been adequately
reported a significant decline in motor evoked potentials when they studied. Boggio et al. (2008) found that the OCC–F8 montage (i.e. cath-
compared left M1-contralateral EC with left M1-contralateral forehead odal stimulation at F8, used as an active control group) was still able to
simulations. It is not known whether such a difference is of clinical induce a larger effect than sham tDCS, but results from this trial were in-
significance, though a pilot study of the F3–EC montage suggests it may conclusive due to the possible confounding effects of concurrent anodal
have more rapid antidepressant effects than the F3–F8 montage. This is occipital stimulation and a small sample size.
presumably due to more effective stimulation of other brain regions Brain regions other than the DLPFC have also been implicated in
(Martin et al., 2011), though the relative efficacy of these approaches depression. These include bilateral sgACC, NAcc, VS/VC, and other
has yet to be tested in an adequately powered randomised controlled central structures, with evidence to suggest that specific aspects of
S. Bai et al. / NeuroImage 87 (2014) 332–344 341

F3-F4-1
Whole brain
F3-F8
F3-EC
Post
L hemisphere Ant
Lat
Med
R hemisphere

CB

L DLPFC

R DLPFC

L OFC

R OFC

L ACC

R ACC

L hippocampus

R hippocampus
Unit: V/m
0 0.04 0.08 0.12 -0.01 0 0.01

Fig. 8. The sensitivity of average E-field magnitude E in brain regions of interest using Msub-aniso, with displacements of F3 electrode with F3–F4-1, F3–F8 and F3–EC. The bar charts on the
left represent the average E-field magnitude when F3 is at the original position. The data points on the right represent the absolute differences in average E-fields from the original
placement.

depressive symptomatology may be linked to particular brain regions Theoretically, it is possible for electrode montages to be tailored to
(Anderson et al., 2012). For example, cognitive aspects (pathological an individual patient's depressive symptoms. For example, montages
guilt, impaired concentration, suicidal ideation) may be linked to ACC, that result in greater stimulation of deep central structures (F3–EC,
OFC and PFC dysfunction, whereas anhedonia, depressed mood and SO–OCC, SO–CB, SO–EC) may be effective for treating melancholic
anxiety may be mediated by the NAcc, sgACC and OFC. Central struc- depression. However, stimulation of the same central structures may
tures including the hypothalamus, locus coeruleus and dorsal raphe increase the risk of a manic episode in bipolar depression (Gálvez
nuclei may be responsible for neurovegetative symptoms such as et al., 2011). In addition to montages used in clinical practice, several hy-
sleep and appetite disturbances. At present, the relative importance of pothetical montages were also modelled. Though the SO–OCC montage
stimulating such regions remains unclear and it is also unknown wheth- was designed for maximal stimulation of ACC and central structures
er frontal montages adopted since 2000 are more advantageous in (thalamus, striatum), our modelling results suggest that ACC modula-
treating depression than the prevailing bilateral SO–EC montage used tion is achieved as readily with the F3–EC montage. The mean E-fields
in earlier decades, which results in strong ACC and OFC activation. The in the DLPFC are also similar for the two montages. Given the promising
diffuse nature of tDCS may be therapeutically advantageous, given the results in an open label depression trial of the F3–EC montage (Martin
widespread anatomy of areas involved in depression, as all of the afore- et al., 2011), a similar antidepressant efficacy is also expected from the
mentioned areas are stimulated to some extent. This is particularly the SO–OCC montage. High E-fields were found in the cerebellum and hip-
case for montages with an extracephalic or posterior electrode, which pocampus with the SO–CB and TMP–EC montages respectively, and
have been shown to lead to more widespread brain current distribution these montages may be useful where modulation of these brain regions
than montages with a contralateral cephalic cathode. is particularly desired. Both cerebellar and temporal lobe changes have
342 S. Bai et al. / NeuroImage 87 (2014) 332–344

been reported in depression (Fitzgerald et al., 2008; Hoppenbrouwers regarding the effects of these montages, simulations using head models
et al., 2008), and it is possible that neuromodulation targeted at these from more subjects with the same electrical properties may be neces-
areas may be beneficial in treating depression. sary as to further substantiate our results.
The E-field directions provide a preliminary understanding of the Nevertheless, the simulation results presented in this study are in
current pathways within the brain, since the current density vector in line with other modelling studies. Parazzini et al. (2011) reported a
a conductive medium is in the same direction as the E-field vector. median E-field magnitude of 0.33 V/m in the cortex using the left M1–
Local E-field vectors generally follow the dominant directionality of right SO configuration, about five times the average E-field value
the E-field in the entire model, which is from anode to cathode. Howev- found here. This is probably due to the variation in head geometry, as
er, the E-field vectors may not all be in parallel with the main E-field well as the adoption of a three-layer skull in our head model, and the
direction, depending on the location of the ROI relative to the elec- difference in the tissue conductivities chosen, which likely resulted in
trodes, and on the local electrical properties. For instance, WM fibres the disparities in the brain E-field magnitude. In addition, Bikson et al.
provide a preferential pathway for current flow along the fibres when (2010) found a peak cortical E-field value of 0.44 V/m in simulations
the flow is partially aligned with the fibre orientation, whereas higher of a healthy subject with left M1–contralateral forehead tDCS. Datta
E-field magnitudes are generated when current flow is transverse to et al. (2011) simulated the head of a stroke patient, and found a peak
the fibre orientation (Bai et al., 2013; Lee et al., 2012). Since the modu- value at 0.36 V/m with the left M1–contralateral shoulder montage.
lation of neurons may also be dependent on directionality of the E-field Those studies, however, assumed a higher value of skull conductivity
(Basser and Roth, 2000; Creutzfeldt et al., 1962; Molaee-Ardekani et al., than the three-layer skull used in this study.
2012), including directionality in analysis may provide more compre-
hensive information on brain activation, though the clinical implications Conclusion
are as yet unclear.
While changes in the electrode size or orientation resulted in differ- With the aid of computational modelling, this study presented a
ences in the E-field magnitude, the differences were not large. This was systematic analysis of various tDCS electrode montages that have been
perhaps due to a relatively small difference in electrode sizes (35 cm2 used in clinical trials, as well as hypothetical montages specifically for
vs. 16 cm2), indeed previous studies have shown that large differences the treatment of depression. The overall profile of E-field magnitude
in E-field magnitude are only found when there is a considerable change and direction, as well as the average E-field magnitudes in key brain
in electrode size (Miranda et al., 2009; Parazzini et al., 2011). Similarly, regions were presented. While the clinical significance of these
displacing the electrode by 1 cm around the actual position did not outcomes is not as yet well understood, these results may form a useful
result in a large difference. These results suggest that electrodes may reference for researchers seeking to optimise the therapeutic potential
be adjusted slightly to accommodate other priorities, such as skin dis- of tDCS by manipulation of electrode montage. Sensitivity analysis on
comfort and damage. For example, rotation of an electrode to avoid the effects of changing the size of the anode, rotating both electrodes
stimulation over a skin lesion such as a mole, may not substantially and displacing the anode was also conducted on selected montages.
affect neuronal outcomes. This is the first computational study dedicated to this type of analysis.
An interesting finding from our results is that current density was This approach could equally inform the use of tDCS to treat other neuro-
not evenly spread over the cerebral cortex, but that local hot spots psychiatric disorders, or the design of montages for neuropsychological
were evident. This phenomenon has been previously reported in experiments.
modelling studies (Datta et al., 2009; Salvador et al., 2010) and indicate
that these localised areas of intense activation may be of clinical impor- Acknowledgment
tance for both efficacy and safety. It appears that the location of the hot
spot is determined by individual anatomical variations in cerebral folds, The authors would like to thank Prof. Caroline Rae and Dr. John Geng
thickness of the CSF layer, skull and scalp, and may contribute to inter- from Neuroscience Research Australia for their support in acquiring and
individual differences in efficacy. processing the structural MRI and DT-MRI data, Dr. Elizabeth Tancred
from the University of New South Wales for her expertise in the head
Model validation and limitations anatomy, and Dr. Angelo Alonzo from the Black Dog Institute for his
help in determining the accurate position of scalp electrodes.
The novelty of this study is that the anatomically-accurate head
models from two different subjects were employed, each with three-
Appendix A. Supplementary data
layered skull structure and one also with WM anisotropy, in order to
investigate several tDCS montages specifically targeted at depression
Supplementary data to this article can be found online at http://dx.
treatment. This was achieved by comparing E-fields in several brain
doi.org/10.1016/j.neuroimage.2013.11.015.
ROIs. Even though some of the above modelling features have appeared
in other studies, they have not been combined together in a single study
References
as has been done in this work. In addition, there has been no modelling
study dedicated to examining tDCS electrode montages which may be Akhtari, M., Bryant, H., Mamelak, A., Heller, L., Shih, J., Mandelkern, M., Matlachov, A.,
useful in the treatment of depression. Ranken, D., Best, E., Sutherling, W., 2000. Conductivities of three-layer human skull.
Brain Topogr. 13, 29–42.
As noted by Lee et al. (2012), validation of computer simulations of Akhtari, M., Bryant, H.C., Mamelak, A.N., Flynn, E.R., Heller, L., Shih, J.J., Mandelkem,
brain E-field distribution against experimental data remains a chal- M., Matlachov, A., Ranken, D.M., Best, E.D., DiMauro, M.A., Lee, R.R., Sutherling,
lenge. As our head model was only comprised of a limited number of W.W., 2002. Conductivities of three-layer live human skull. Brain Topogr. 14,
151–167.
tissues, errors may arise from the exclusion of the dielectric properties Anderson, R., Frye, M.A., Abulseoud, O.A., Lee, K.H., McGillivray, J., Berk, M., Tye, S.J., 2012.
of other tissues, even though their electrical conductivities were of the Deep brain stimulation for treatment-resistant depression: efficacy, safety and mech-
same order as that of the substitute compartment. The conductivities anisms of action. Neurosci. Biobehav. Rev. 36, 1920–1933.
Arul-Anandam, A., Loo, C., 2009. Transcranial direct current stimulation: a new tool for
were chosen as an average across different experimental studies,
the treatment of depression? J. Affect. Disord. 117, 137–145.
which suggested that they were not measured with the same experi- Bai, S., Loo, C., Dokos, S., 2013. A review of computational models of transcranial electrical
mental setup. This, together with the artificial approximation of the stimulation. Crit. Rev. Biomed. Eng. 41, 21–35.
torso, may contribute to computational errors. As DTI was not per- Basser, P., Roth, B., 2000. New currents in electrical stimulation of excitable tissues. Annu.
Rev. Biomed. Eng. 2, 377–397.
formed on Fsub, Msub also served as a transition model. However, in Baumann, S., Wozny, D., Kelly, S., Meno, F., 1997. The electrical conductivity of human
order to provide a more generalised conclusion for clinical investigation cerebrospinal fluid at body temperature. IEEE Trans. Biomed. Eng. 44, 220–223.
S. Bai et al. / NeuroImage 87 (2014) 332–344 343

Behrens, T., Johansen-Berg, H., Woolrich, M., Smith, S., Wheeler-Kingshott, C., Boulby, P., Gonçalves, S., de Munck, J., Verbunt, J., Bijma, F., Heethaar, R., Lopes da Silva, F., 2003. In
Barker, G., Sillery, E., Sheehan, K., Ciccarelli, O., Thompson, A., Brady, J., Matthews, vivo measurement of the brain and skull resistivities using an EIT-based method
P., 2003a. Non-invasive mapping of connections between human thalamus and and realistic models for the head. IEEE Trans. Biomed. Eng. 50, 754–767.
cortex using diffusion imaging. Nat. Neurosci. 6, 750–757. Greicius, M., Flores, B., Menon, V., Glover, G., Solvason, H., Kenna, H., Reiss, A., Schatzberg,
Behrens, T., Woolrich, M., Jenkinson, M., Johansen-Berg, H., Nunes, R., Clare, S., Matthews, A., 2007. Resting-state functional connectivity in major depression: abnormally in-
P., Brady, J., Smith, S., 2003b. Characterization and propagation of uncertainty in creased contributions from subgenual cingulate cortex and thalamus. Biol. Psychiatry
diffusion-weighted MR imaging. Magn. Reson. Med. 50, 1077–1088. 62, 429–437.
Behrens, T., Berg, H., Jbabdi, S., Rushworth, M., Woolrich, M., 2007. Probabilistic diffusion Grimm, S., Beck, J., Schuepbach, D., Hell, D., Boesiger, P., Bermpohl, F., Niehaus, L.,
tractography with multiple fibre orientations: what can we gain? NeuroImage 34, Boeker, H., Northoff, G., 2008. Imbalance between left and right dorsolateral
144–155. prefrontal cortex in major depression is linked to negative emotional judg-
Bewernick, B., Hurlemann, R., Matusch, A., Kayser, S., Grubert, C., Hadrysiewicz, B., ment: an fMRI study in severe major depressive disorder. Biol. Psychiatry 63,
Axmacher, N., Lemke, M., Cooper-Mahkorn, D., Cohen, M., Brockmann, H., Lenartz, 369–376.
D., Sturm, V., Schlaepfer, T., 2010. Nucleus accumbens deep brain stimulation Gutierrez, D., Nehorai, A., Muravchik, C., 2004. Estimating brain conductivities and dipole
decreases ratings of depression and anxiety in treatment-resistant depression. Biol. source signals with EEG arrays. IEEE Trans. Biomed. Eng. 51, 2113–2122.
Psychiatry 67, 110–116. Hoppenbrouwers, S., Schutter, D., Fitzgerald, P., Chen, R., Daskalakis, Z., 2008. The role of
Bewernick, B., Kayser, S., Sturm, V., Schlaepfer, T., 2012. Long-term effects of nucleus the cerebellum in the pathophysiology and treatment of neuropsychiatric disorders:
accumbens deep brain stimulation in treatment-resistant depression: evidence for a review. Brain Res. Rev. 59, 185–200.
sustained efficacy. Neuropsychopharmacology 37, 1975–1985. Johansen-Berg, H., Behrens, T., 2009. Diffusion MRI: From Quantitative Measurement to
Bikson, M., Datta, A., Rahman, A., Scaturro, J., 2010. Electrode montages for tDCS and weak In-vivo Neuroanatomy. Academic Press, London.
transcranial electrical stimulation: role of “return” electrode's position and size. Clin. Kalu, U., Sexton, C., Loo, C., Ebmeier, K., 2012. Transcranial direct current stimulation in
Neurophysiol. 121, 1976–1978. the treatment of major depression: a meta-analysis. Psychol. Med. 1, 1–10.
Boggio, P., Rigonatti, S., Ribeiro, R., Myczkowski, M., Nitsche, M., Pascual-Leone, A., Fregni, Lai, Y., Van Drongelen, W., Ding, L., Hecox, K., Towle, V., Frim, D., He, B., 2005. Estima-
F., 2008. A randomized, double-blind clinical trial on the efficacy of cortical direct cur- tion of in vivo human brain-to-skull conductivity ratio from simultaneous extra-
rent stimulation for the treatment of major depression. Int. J. Neuropsychopharmacol. and intra-cranial electrical potential recordings. Clin. Neurophysiol. 116,
11, 249–254. 456–465.
Borckardt, J., Bikson, M., Frohman, H., Reeves, S., Datta, A., Bansal, V., Madan, A., Barth, Lee, W., Deng, Z., Kim, T., Laine, A., Lisanby, S., Peterchev, A., 2012. Regional electric field
K., George, M., 2012. A pilot study of the tolerability and effects of high-definition induced by electroconvulsive therapy in a realistic finite element head model: influ-
transcranial direct current stimulation (HD-tDCS) on pain perception. J. Pain 13, ence of white matter anisotropic conductivity. NeuroImage 59, 2110–2123.
112–120. Lippold, O., Redfearn, J., 1964. Mental changes resulting from the passage of small direct
Brunoni, A., Ferrucci, R., Bortolomasi, M., Vergari, M., Tadini, L., Boggio, P., Giacopuzzi, M., currents through the human brain. Br. J. Psychiatry 110, 768–772.
Barbieri, S., Priori, A., 2011. Transcranial direct current stimulation (tDCS) in unipolar Loo, C., Sachdev, P., Martin, D., Pigot, M., Alonzo, A., Malhi, G., Lagopoulos, J., Mitchell, P.,
vs. bipolar depressive disorder. Prog. Neuropsychopharmacol. Biol. Psychiatry 35, 2010. A double-blind, sham-controlled trial of transcranial direct current stimulation
96–101. for the treatment of depression. Int. J. Neuropsychopharmacol. 13, 61–69.
Brunoni, A., Valiengo, L., Baccaro, A., Zanão, T., de Oliveira, J., Goulart, A., Boggio, P., Lotufo, Loo, C., Alonzo, A., Martin, D., Mitchell, P., Galvez, V., Sachdev, P., 2012. Transcranial direct
P., Benseñor, I., Fregni, F., 2013. The sertraline vs electrical current therapy for treating current stimulation for depression: 3-week, randomised, sham-controlled trial. Br.
depression clinical study: results from a factorial, randomized, controlled trial. JAMA J. Psychiatry 200, 52–59.
Psychiatry 70, 383–391. Lozano, A., Giacobbe, P., Hamani, C., Rizvi, S., Kennedy, S., Kolivakis, T., Debonnel, G.,
Carballedo, A., Scheuerecker, J., Meisenzahl, E., Schoepf, V., Bokde, A., Möller, H., Doyle, M., Sadikot, A., Lam, R., Howard, A., Ilcewicz-Klimek, M., Honey, C., Mayberg, H., 2012.
Wiesmann, M., Frodl, T., 2011. Functional connectivity of emotional processing in A multicenter pilot study of subcallosal cingulate area deep brain stimulation for
depression. J. Affect. Disord. 134, 272–279. treatment-resistant depression. J. Neurosurg. 116, 315–322.
Creutzfeldt, O., Fromm, G., Kapp, H., 1962. Influence of transcortical DC currents on corti- Malone, D., Dougherty, D., Rezai, A., Carpenter, L., Friehs, G., Eskandar, E., Rauch, S.,
cal neuronal activity. Exp. Neurol. 5, 436–452. Rasmussen, S., Machado, A., Kubu, C., Tyrka, A., Price, L., Stypulkowski, P., Giftakis, J.,
Datta, A., Elwassif, M., Battaglia, F., Bikson, M., 2008. Transcranial current stimula- Rise, M., Malloy, P., Salloway, S., Greenberg, B., 2009. Deep brain stimulation of the
tion focality using disc and ring electrode configurations. J. Neural Eng. 5, ventral capsule/ventral striatum for treatment-resistant depression. Biol. Psychiatry
163–174. 65, 267–275.
Datta, A., Bansal, V., Diaz, J., Patel, J., Reato, D., Bikson, M., 2009. Gyri-precise head model of Martin, D., Alonzo, A., Mitchell, P., Sachdev, P., Gálvez, V., Loo, C., 2011. Fronto-
transcranial direct current stimulation: improved spatial focality using a ring elec- extracephalic transcranial direct current stimulation as a treatment for major depres-
trode versus conventional rectangular pad. Brain Stimul. 2, 201–207. sion: an open-label pilot study. J. Affect. Disord. 134, 459–463.
Datta, A., Baker, J.M., Bikson, M., Fridriksson, J., 2011. Individualized model predicts brain Mayberg, H., 1997. Limbic–cortical dysregulation: a proposed model of depression.
current flow during transcranial direct-current stimulation treatment in responsive J. Neuropsychiatry Clin. Neurosci. 9, 471–481.
stroke patient. Brain Stimul. 4, 169–174. Mayberg, H., 2003. Modulating dysfunctional limbic–cortical circuits in depression:
Datta, A., Dmochowski, J., Guleyupoglu, B., Bikson, M., Fregni, F., 2012. Cranial electrother- towards development of brain-based algorithms for diagnosis and optimised treat-
apy stimulation and transcranial pulsed current stimulation: a computer based high- ment. Br. Med. Bull. 65, 193–207.
resolution modeling study. NeuroImage 65, 280–287. Mayberg, H., 2007. Defining the neural circuitry of depression: toward a new nosology
Dell'Osso, B., Zanoni, S., Ferrucci, R., Vergari, M., Castellano, F., D'Urso, N., Dobrea, C., with therapeutic implications. Biol. Psychiatry 61, 729–730.
Benatti, B., Arici, C., Priori, A., Altamura, A., 2012. Transcranial direct current stimula- Mayberg, H., Lozano, A., Voon, V., McNeely, H., Seminowicz, D., Hamani, C., Schwalb, J.,
tion for the outpatient treatment of poor-responder depressed patients. Eur. Psychi- Kennedy, S., 2005. Deep brain stimulation for treatment-resistant depression. Neuron
atry 27, 513–517. 45, 651–660.
Dreyfuss, H., Tilley, A., 1993. Measure of Man and Woman: Human Factors in Design. John Mendonca, M., Santana, M., Baptista, A., Datta, A., Bikson, M., Fregni, F., Araujo, C., 2011.
Wiley & Sons, NY. Transcranial dc stimulation in fibromyalgia: optimized cortical target supported by
Ferrucci, R., Bortolomasi, M., Brunoni, A., Vergari, M., Tadini, L., Giacopuzzi, M., Priori, A., high-resolution computational models. J. Pain 12, 610–617.
2009a. Comparative benefits of transcranial direct current stimulation (tDCS) treat- Miranda, P., Lomarev, M., Hallett, M., 2006. Modeling the current distribution during
ment in patients with mild/moderate vs. severe depression. Clin. Neuropsychiatry transcranial direct current stimulation. Clin. Neurophysiol. 117, 1623–1629.
6, 246–251. Miranda, P., Faria, P., Hallett, M., 2009. What does the ratio of injected current to electrode
Ferrucci, R., Bortolomasi, M., Vergari, M., Tadini, L., Salvoro, B., Giacopuzzi, M., Barbieri, S., area tell us about current density in the brain during tDCS? Clin. Neurophysiol. 120,
Priori, A., 2009b. Transcranial direct current stimulation in severe, drug-resistant 1183–1187.
major depression. J. Affect. Disord. 118, 215–219. Molaee-Ardekani, B., Márquez-Ruiz, J., Merlet, I., Leal-Campanario, R., Gruart, A., Sánchez-
Fitzgerald, P., Laird, A., Maller, J., Daskalakis, Z., 2008. A meta-analytic study of changes in Campusano, R., Birot, G., Ruffini, G., Delgado-García, J., Wendling, F., 2012. Effects of
brain activation in depression. Hum. Brain Mapp. 29, 683–695. transcranial direct current stimulation (tDCS) on cortical activity: a computational
Fox, M., Buckner, R., White, M., Greicius, M., Pascual-Leone, A., 2012. Efficacy of transcra- modeling study. Brain Stimul. http://dx.doi.org/10.1016/j.brs.2011.12.006.
nial magnetic stimulation targets for depression is related to intrinsic functional con- Moliadze, V., Antal, A., Paulus, W., 2010. Electrode-distance dependent after-effects of
nectivity with the subgenual cingulate. Biol. Psychiatry 72, 595–603. transcranial direct and random noise stimulation with extracephalic reference elec-
Fregni, F., Boggio, P., Nitsche, M., Marcolin, M., Rigonatti, S., Pascual-Leone, A., 2006a. trodes. Clin. Neurophysiol. 121, 2165–2171.
Treatment of major depression with transcranial direct current stimulation. Bipolar Nicholson, P., 1965. Specific impedance of cerebral white matter. Exp. Neurol. 13,
Disord. 8, 203–204. 386–401.
Fregni, F., Boggio, P., Nitsche, M., Rigonatti, S., Pascual-Leone, A., 2006b. Cognitive effects Nitsche, M., Paulus, W., 2000. Excitability changes induced in the human motor cortex by
of repeated sessions of transcranial direct current stimulation in patients with weak transcranial direct current stimulation. J. Physiol. 527, 633–639.
depression. Depress. Anxiety 23, 482–484. Nitsche, M., Paulus, W., 2001. Sustained excitability elevations induced by transcranial DC
Gabriel, C., Gabriel, S., Corthout, E., 1996. The dielectric properties of biological tissues: I. motor cortex stimulation in humans. Neurology 57, 1899–1901.
Literature survey. Phys. Med. Biol. 41, 2231–2249. Oostendorp, T., Delbeke, J., Stegeman, D., 2000. The conductivity of the human skull:
Gálvez, V., Alonzo, A., Martin, D., Mitchell, P.B., Sachdev, P., Loo, C.K., 2011. Hypomania results of in vivo and in vitro measurements. IEEE Trans. Biomed. Eng. 47,
induction in a patient with bipolar II disorder by transcranial direct current stimula- 1487–1492.
tion (tDCS). J. ECT 27, 256–258. Palm, U., Schiller, C., Fintescu, Z., Obermeier, M., Keeser, D., Reisinger, E., Pogarell, O.,
Geddes, L., Baker, L., 1967. The specific resistance of biological material — a com- Nitsche, M., Möller, H., Padberg, F., 2011. Transcranial direct current stimulation in
pendium of data for the biomedical engineer and physiologist. Med. Biol. Eng. treatment resistant depression: a randomized double-blind, placebo-controlled
5, 271–293. study. Brain Stimul. http://dx.doi.org/10.1016/j.brs.2011.08.005.
344 S. Bai et al. / NeuroImage 87 (2014) 332–344

Parazzini, M., Fiocchi, S., Rossi, E., Paglialonga, A., Ravazzani, P., 2011. Transcranial direct Shattuck, D., Leahy, R., 2002. BrainSuite: an automated cortical surface identification tool.
current stimulation: estimation of the electric field and of the current density in an Med. Image Anal. 6, 129–142.
anatomical human head model. IEEE Trans. Biomed. Eng. 58, 1773–1780. Suh, H.S., Lee, W.H., Kim, T.S., 2012. Influence of anisotropic conductivity in the skull and
Pittenger, C., Duman, R.S., 2007. Stress, depression, and neuroplasticity: a convergence of white matter on transcranial direct current stimulation via an anatomically realistic
mechanisms. Neuropsychopharmacology 33, 88–109. finite element head model. Phys. Med. Biol. 57, 6961–6980.
Redfearn, J., Lippold, O., Costain, R., 1964. Preliminary account of the clinical effects of Wagner, T., Fregni, F., Fecteau, S., Grodzinsky, A., Zahn, M., Pascual-Leone, A., 2007. Trans-
polarizing the brain in certain psychiatric disorders. Br. J. Psychiatry 110, cranial direct current stimulation: a computer-based human model study.
773–785. NeuroImage 35, 1113–1124.
Salvador, R., Mekonnen, A., Ruffini, G., Miranda, P., 2010. Modeling the electric field Weaver, L., Williams, R., Rush, S., 1976. Current density in bilateral and unilateral ECT.
induced in a high resolution realistic head model during transcranial current stimu- Biol. Psychiatry 11, 303–312.
lation. Conf Proc IEEE Eng Med Biol Soc, pp. 2073–2076. Wolters, C., Anwander, A., Tricoche, X., Weinstein, D., Koch, M., MacLeod, R., 2006. Influ-
Shahid, S., Wen, P., Ahfock, T., 2013. Numerical investigation of white matter anisotropic ence of tissue conductivity anisotropy on EEG/MEG field and return current compu-
conductivity in defining current distribution under tDCS. Comput. Methods Programs tation in a realistic head model: a simulation and visualization study using high-
Biomed. 109, 48–64. resolution finite element modeling. NeuroImage 30, 813–826.
Reproduced with permission of the copyright owner. Further reproduction prohibited without
permission.

S-ar putea să vă placă și