Вы находитесь на странице: 1из 14

www.elsevier.

com/locate/ynimg
NeuroImage 25 (2005) 369 – 382

Evaluation of different cortical source localization methods using


simulated and experimental EEG data
Jun Yaoa and Julius P.A. Dewalda,b,c,d,*
a
Department of Physical Therapy and Human Movement Sciences, Northwestern University, Chicago, IL 60611, USA
b
Department of Physical Med and Rehab., Northwestern University, Chicago, IL 60611, USA
c
Department of Biomedical Engineering, Northwestern University, Evanston, IL 60201, USA
d
Sensory Motor Performance Program, Rehabilitation Institute of Chicago, Chicago, IL 60611, USA

Received 8 February 2004; revised 23 July 2004; accepted 29 November 2004

Different cortical source localization methods have been developed to measurements of cortical activity with a millisecond temporal
directly link the scalp potentials with the cortical activities. Up to now, resolution. However, the spatial resolution of EEG recordings is
these methods are the only possible solution to noninvasively inves- limited, which is mainly caused by the blurring effect of the head
tigate cortical activities with both high spatial and time resolutions. volume conductor. In addition, the effects of electrode density, i.e.,
However, the application of these methods is hindered by the fact that spatial sampling, signal-to-noise ratio (SNR) of the recordings, and
they have not been rigorously evaluated nor compared. In this paper,
the reference electrode placement, all contribute to the level of
the performances of several source localization methods (moving
dipoles, minimum Lp norm, and low resolution tomography (LRT) spatial resolution obtained by inverse calculations with EEG
with Lp norm, p equal to 1, 1.5, and 2) were evaluated by using (Nunez, 1981). To improve the spatial resolution of scalp EEG,
simulated scalp EEG data, scalp somatosensory evoked potentials tremendous efforts have been made to use cortical source local-
(SEPs), and upper limb motor-related potentials (MRPs) obtained on ization techniques to de-blur the scalp EEG recordings by de-
human subjects (all with 163 scalp electrodes). By using simulated EEG convolving the measured scalp EEG distribution into an electrical
data, we first evaluated the source localization ability of the above activity (current or potential) over the cortical surface. These
methods quantitatively. Subsequently, the performance of the various techniques are expected to improve the spatial resolution from
methods was evaluated qualitatively by using experimental SEPs and centimeter scale on the scalp to millimeter scale on the cortex, and
MRPs. Our results show that the overall LRT Lp norm method with p thus make it possible to show cortical spatiotemporal activities
equal to 1 has a better source localization ability than any of the other
with a significant improvement in spatial resolution in a non-
investigated methods and provides physiologically meaningful recon-
struction results. Our evaluation results provide useful information for invasive way.
choosing cortical source localization approaches for future EEG/MEG In order to de-convolve the scalp EEG, a head volume con-
studies. ductor model and a source model are required. As for the head
D 2004 Elsevier Inc. All rights reserved. volume conductor model, a homogeneous sphere head volume
conductor model, a boundary element method (BEM) model, or a
Keywords: Brain imaging; Source localization; Dipole fit; Current density finite element method (FEM) model can be used. Since the BEM
reconstruction; Somatosensory evoked potentials; Event-related potentials model is a compromise between oversimplifying sphere head
volume model and the large computational costs of a FEM model,
the subject specific BEM model is currently most widely used
(Fuchs et al., 1998). As for the source model, two commonly used
Introduction models are equivalent current dipole models (Scherg, 1990; Scherg
and Buchner, 1993; Williamson and Kaufman, 1981) and current
Brain activity has both spatial and temporal characteristics. distributed source models (Hamalainen and Ilmoniemi, 1994;
Conventional scalp EEG recordings are noninvasive global Ilmoniemi, 1993; Nenonen et al., 1994; Pascual-Marqui and
Biscay-Lirio, 1993; Pascual-Marqui et al., 1994; Wagner et al.,
2000; Wang et al., 1992, 1993).
* Corresponding author. Department of Physical Therapy and Human
Movement Sciences, Biomedical Engineering, Northwestern University, When using any source models, regularized solutions are
645 N. Michigan Ave, Suite 1100, Room 1159, Chicago, IL 60611, USA. commonly used to solve the ill-posed inverse problem. With the
Fax: +1 312 908 0741. combination of different head models, source models, and
E-mail address: j-dewald@northwestern.edu (J.P.A. Dewald). associated regulation methods, there are currently dozens of
Available online on ScienceDirect (www.sciencedirect.com). different cortical source localization methods available. The
1053-8119/$ - see front matter D 2004 Elsevier Inc. All rights reserved.
doi:10.1016/j.neuroimage.2004.11.036
370 J. Yao, J.P.A. Dewald / NeuroImage 25 (2005) 369–382

advantages and disadvantages of these methods were analyzed in (For the convenience of representation, in the later parts of this
principle and described qualitatively by Wagner et al. (2000). paper, we call these three simulated EEG data sets as dsimulated
Comparisons between the ability of methods to detect source data sets I/II/IIIT.) A single dipole area (simulated data set I) was
location(s) have been reported in the literature using experimental used to simulate EEG data generated from a concentrated source.
EEG data (Cincotti et al., 2004; Haan et al., 2000; Komssi et al., Two equal-size areas of dipoles with increasing area size and two
2004; Pascual-Marqui et al., 1994; Phillips et al., 2002; Stenbacka dipoles with increasing distance (simulated data sets II and III) were
et al., 2002; Waberski et al., 2000) and simulated EEG data used to simulate EEG data generated from multiple sources.
(Babiloni et al., 2000b, 2004; Baillet and Garnero, 1997; He et al., Each EEG data set is comprised of a total of 163 EEG signals,
2002a) with different number of EEG electrodes (ranging from 19 which are the forward calculation results obtained at the electrodes
to 128), head models, and regulation parameters involved. All on the scalp with a reference at Oz. The forward calculation was
these previous studies incorporated comparisons of two or at most performed using a realistic BEM model, which consists of three
three methods with fixed regulation settings. The purpose of this shells: skin, skull, and liquor. The BEM model is developed based
paper is to expand comparisons to a large number of source on the MRI data (with a resolution of 1.0  1.0  1.0 mm) of a
reconstruction methods, including dipole methods (single and dual normal subject. The parameters of BEM model are listed in Table 1.
moving dipole), minimum Lp norm methods ( p equals to 1, 1.5, Since real EEG recordings are usually noisy, we further added
and 2), and low resolution electromagnetic brain tomography noise to the simulated EEG data. Noise was obtained from a real
(LRT) methods with various Lp norms ( p equals to 1, 1.5, and 2). EEG measurement in which the subject was not performing any
In an effort to estimate the accuracy of these source localization task and with his eyes open and looking at a fixed point. One
methods, we first created simulated EEG data with the obvious hundred epochs were averaged to achieve the noise levels
advantage that the exact source locations/regions were generated comparable to the typical event-related potentials (ERPs). This
by the investigator. The shortcomings of using simulated EEG data corresponds to a SNR of approximately 9–10. Subsequently, the
are (1) the data are not realistic and (2) since the forward and data with noise were low-pass filtered by a 9th order zero-lag filter
backward calculations are based on the same head model, the with the cut-off frequency of 50 Hz.
results obtained based on simulated EEG cannot reflect errors
introduced by the head model but only show the accuracy of source Experimental EEG data
model and associated regulation methods. In addition to using the
simulated EEG data, we also qualitatively tested the performance Each of the inverse methods was also applied to two types of
of the various source localization methods using experimental EEG experimental EEG data, i.e., somatosensory evoked potentials
data, including somatosensory evoked potentials (SEPs) from (SEPs) and event-related potentials (ERPs).
electrical stimulation of the wrist, elbow, and shoulder, and event-
related potentials (ERPs) resulting from the generation of isometric Recording and processing of somatosensory evoked potentials
shoulder abduction torques. The advantage of using SEPs and SEPs resulting from electrical stimulation of right wrist, elbow,
ERPs is that the location of sensory-motor cortical activity is well or shoulder of one subject were recorded by a BioSemi Active II
described in the literature ((Babiloni et al., 2000a; Druschky et al., system (BioSemi, Amsterdam, Netherlands) with 163 active
2003; Emerson and Pedley, 1984; Hari and Forss, 1999; electrodes and a reference on the right ear lobe. Constant current
Hashimoto, 2000; He et al., 2002b; Recuero, 1999; Sonoo, 2000; square-wave pulses of 0.3 ms duration were delivered to the
Towle et al., 2003; Valeriani et al., 2000) for SEPs and (Penfield median nerve at the wrist, or lateral epicondyle elbow, or the
and Rasmussen, 1950; Rao et al., 1995) for ERPs), which acromion of the shoulder through disposable neurology electrodes
permitted us to determine whether solutions obtained with the (Ambu USA, Linthicum, MD) with 1 cm between anode and
various source localization methods were reasonable or not. cathode using a Compex2 system (Compex, Annecy-le-Vieux,
In the following parts of this paper, a BEM realistic head model France). An interstimulus interval of 500 ms was used for a total of
was used for both EEG data simulation and inverse calculation. All 1000 stimuli for each stimulus location. The intensity of the pulses
the computations were performed using the Curry V4.5 (Neuroscan, was set to the highest amplitude that the subject could tolerate
TX) software package, which is a widely used commercial source without discomfort. The SEP data were sampled at 2 kHz. The
localization software package. Other source localization methods, positions of each EEG electrode and three fiducial landmarks
such as cortical potential imaging method (He et al., 2002b) and the (Nasion, left preauricular point and right preauricular point) were
three-dimensional resultant vector method (Ricamato et al., 2003), marked by a Polhemus (Polhemus, Colchester, VT) three-dimen-
are not included in this study. Parts of this work have been sional magnetic digitizer. The three fiducial landmarks were used
published as a conference preceding (Yao and Dewald, 2003). to co-register the 163 EEG electrodes with the subject’s MRI data.
The co-registration was repeated by adjusting the fiducial land-
marks in MRI data until all the 163 electrodes were co-registered
Materials and methods well with the subject’s scalp.

Simulation EEG data


Table 1
Parameters of the BEM model
By placing equal strength (1 AAmm) dipoles on the cortex, we
simulated three EEG data sets, i.e., (I) single dipole area with a fixed Shell Surface Side length Conductivity
triangle no. (mm) (S/m)
center and an increasing radius ranging from 5 mm to 10 mm, (II)
two equal-size areas of dipoles with fixed centers and increasing Skin 2692 10.0 0.25
radiuses ranging from 5 mm to 10 mm, and (III) two dipoles with Skull 2418 9.0 0.017
increasing distance between them ranging from 4 mm to 20 mm. Liquor 3306 7.0 1.79
J. Yao, J.P.A. Dewald / NeuroImage 25 (2005) 369–382 371

The SEPs were read into Brain Vision Analyzer V1.04 software provided during the data collection section; instead, the subject
(Brain Products GmbH, Mqnchen, Germany), filtered by a 1– was required to look at a fixed point without eye movement
500 Hz (48 dB/oct) band-pass filter with a notch at 60 Hz. Then the during the whole period of a trail (totally about 8 s including 5–6
intervals that contained eye blink signals were eliminated from s before moving and 2 s after moving). Torque generation
further analysis. The data over 40 ms pre- and 60 ms post-stimulus performance was reported to the subject at the end of each trial.
period were baseline corrected and averaged. The EEG data were sampled at 2 kHz. Similarly, co-registration of
EEG electrode positions with the subject’s scalp was performed
Recording and processing of event-related potentials off-line.
Electroencephalographic potentials related to isometric shoul- The event-related potentials (MRPs) were read into Brain
der abduction torque generated by one subject were recorded by Vision Analyzer V1.04 software. The intervals that contained eye
the BioSemi Active II system (BioSemi, Amsterdam, Netherlands). blink signals were eliminated from further analysis. The data over
During the experiment, the subject was casted at the wrist and 2000 ms pre- and 500 ms post-torque onset were baseline corrected
secured to a six-degree-of-freedom (DOF) load cell with the and averaged.
shoulder at a 758 abduction angle and a 408 flexion angle. The tip
of the hand was located at a distance from the body corresponding Source reconstruction
to an elbow angle of 908 with 08 representing full extension of the
elbow. In order to minimize the activation of trunk muscles, A realistic BEM head model and two different source models,
subjects were seated in a Biodex chair with the trunk secured and dipole and distributed current source models, were used to
shoulders strapped to the back of the chair. A monitor was placed reconstruct the cortical source of both simulated and experimental
in front of the subject to provide visual feedback of torque EEG data. Each inverse method was applied to simulated EEG data
generation in the SABD direction. with 10 different time points. Among these 10 points, the noise is
The forces and torques generated at the wrist were measured random. Regularization methods and parameters used in this paper
using a six-degree-of-freedom load cell (JR3 Inc., Woodland, CA) are listed in Fig. 1. The meaning of each of the parameters is
and converted online to torques at the elbow (flexion/extension) described in the Curry reference guide (NeuroScan, 1999). For
and shoulder (flexion/extension, abduction/adduction, and exter- convenience of representation, in the rest of this paper, we refer to
nal/internal rotation) based on a free body analysis of the upper all the source localization methods with different source model,
limb (Beer et al., 1995). Online visual feedback of the SABD different regularization methods, or parameters as methods 1–8.
level was provided during the training section. Maximum The exact makeup of each of the methods is provided by the upper
voluntary torque (MVT) in shoulder abduction direction was first indices in the most right column of Fig. 1 with abbreviations listed
recorded for the subject. The subject was then trained to generate beside the indices.
a self-initiated SABD torque at 25% of his MVT and maintain it Inverse methods using the dipole source model operated on the
for a period of 0.3 s. After the subject felt comfortable with the assumption that only a discrete number of generators (usually
task, EEG data were collected. During the data collection, the dipoles) were active at a given time or over a time period. In this
subject was asked to repeat the same task 100 times in blocks of paper, we investigated single and dual moving dipole methods
20 trials with 10 s rest periods between individual trials and 10 (MDP1 and MDP2). These methods model cortical currents with
min rest periods between blocks to avoid fatigue. In order to equivalent current dipoles that are described by their three-
reduce brain activity at the visual cortex, no visual feedback was dimensional locations, fixed orientation, and variable amplitude

Fig. 1. Settings for the various source localization methods.


372 J. Yao, J.P.A. Dewald / NeuroImage 25 (2005) 369–382

(Williamson and Kaufman, 1981). The position of the dipoles was The percentages of undetected source number (PUS) and of
found using a least-squares search that minimizes the difference falsely-detected source number (PFS) were defined as
between the estimated and measured data. Dipole models require a Nun Nfalse
priori knowledge of the number of sources, which is usually PUS ¼ ; and PFS ¼ ; ð2Þ
Nreal Nestimated
unknown.
Conversely, when dealing with current distributed-source where N un, N false, N real, and N estimated are the numbers of
reconstruction (CDR) methods, no a priori knowledge of the undetected, falsely-detected, real, and estimated sources, respec-
number of sources is required. Since the neurophysiological tively. When calculating PUS and PFS, an undetected source is
potentials follow Ohm’s law, using CDR methods to reconstruct defined as a real source whose location to its closest estimated
the source is a linear underdetermined inverse problem. Solution to source is larger than 0.6 times the unit distance of the source
this problem uses a regulation operator. More specifically, the layer. (Since we constrained all the solutions on a grid of cortex
regulation used in minimum Lp norm methods minimizes the with 5 mm between each point, the unit distance of the source
variance D 2 which reads D 2 = tLx  mtp + ktWxtp , where x is layer is equal to 5 mm.) Similarly, a falsely-detected source is
the vector of source currents; L is the lead field matrix; m is the defined as an estimated source whose location to its closest real
EEG measurements; W is a diagonal location weighting matrix; k source is larger than 0.6 times the unit distance of the source
is the regulation parameter; and p (1 V p V 2) is the measure in layer.
Banach space. When W is equal to identity matrix (as used in this
paper), it is referred to as a regular minimum Lp norm method; Evaluation of the results on real EEG data
otherwise, it is a weighted minimum Lp norm method. By Since the exact locations of SEPs and ERPs are unknown,
changing the weighting matrix W to BW, where B is the Laplacian it is hard to quantitatively evaluate the results obtained by
coupling matrix, it becomes a LORETA (LRT) method. The effect different inverse methods. Qualitative evaluation based on current
of matrix B is to constrain the neighboring sources to similar physiologic knowledge of expected active-region was therefore
strengths and thus smooth the source-current distributions. In this performed.
paper, minimum Lp norm and LRT Lp norm methods with p equal EEG source localization associated with SEP data is relatively
to 1, 1.5, and 2 were investigated. Furthermore, solutions of CDR clean and has been well documented in the literature (Babiloni
methods investigated in this paper were constrained on a cortex et al., 2000a; Emerson and Pedley, 1984; Hashimoto, 2000;
overlay with 5 mm between each point and the directions of Recuero, 1999; Sonoo, 2000; Valeriani et al., 2000). The most
sources were rotated. widely researched and clinical applied SEPs are elicited by
stimulation of the median nerve at the wrist (Emerson and Pedley,
Evaluation methods 1984; Sonoo, 2000; Valeriani et al., 2000). More recently, SEP
data resulting from the stimulation of fingers or other sites have
Evaluation of the results on simulated EEG data also been reported (Dowman and Schell, 1999; Druschky et al.,
The accuracy of the inverse solution using the simulated EEG 2003; Restuccia et al., 2002; Schaefer et al., 2002; Waberski et al.,
data was evaluated by three indices: (1) the error distance (ED), 1999). Different imaging modalities, such as fMRI, PET, and
which is the distance between the locations of real and estimated source localization methods based on EEG/MEG, showed that
sources, (2) the percentage of undetected source number (PUS), Brodmann areas 2 and 3a, which receive deep joint and muscle
and (3) the percentage of falsely-detected source number (PFS). receptor inputs, are believed to be the generator sites for the early
The error distance (ED) between the real and the estimated components of proprioception-related evoked responses, in
source locations were defined as contrast to area 3b activation following cutaneous mechanical
and electrical stimulation, which with clearly somatotopical
1 X
NI   1 XNJ
  representations of different body parts (Druschky et al., 2003;
 min kd˜i  sj k þ  min ksj  dd̃i k : ð1Þ
NI i
j NJ ja J
i Hari and Forss, 1999). Direct recording of cortical potentials in
the human further confirmed the above findings (Towle et al.,
By using this definition, one can measure the averaged error 2003).
between the inverse calculation and the actual source location as a Event-related cortical activity has mostly been studied using
function of distance. In this equation, s i is the real dipole location cortical stimulation or extracellular recordings. Early stimulation
of the simulated EEG data; d̃ i is the source location detected by experiments in monkeys and chimpanzees (Ferrier, 1875; Hines,
each inverse calculation method; i and j are the indices of locations 1940; Sherrington, 1947; Woolsey, 1958) demonstrated that
of estimated and the actual sources; and N I and N J are the total stimulation of the primary motor cortex (M1) resulted in movement
numbers of estimated and the undetected sources, respectively. In in the contralateral side of the body. Penfield and Rasmussen (1950)
the case that CDR methods were used, a source location is defined performed cortical stimulation experiments in humans and reported
as the location where the current strength is larger than the results similar to those found in animal work. Furthermore,
threshold. This threshold is set as the value corresponding to 98% intracortical microstimulation results in the monkey (Asanuma
confident level of a Weibull distribution function (Weibull, 1939) and Rosen, 1972; Schieber, 2001) and from other studies, including
that fits well to the estimated cortical currents. The first term of Eq. single neuron recordings in monkeys (Schieber and Hibbard, 1993),
(1) calculates the mean of the distance from each estimated source fMRI recordings in humans (Sanes et al., 1995), and MEGs in
to its closest real source; and the corresponding real source is then humans (Cheyne et al., 1991; Salenius et al., 1997), all suggested
marked as a detected source. All the undetected real sources made that M1 is the source generator of movement-related potentials.
up the elements of data set J. And the second term of Eq. (1) The previously established knowledge of the source for SEPs
calculates the mean of the distance from each of the undetected and ERPs was used to evaluate the performance of different source
sources to the closest estimated sources. localization methods qualitatively.
J. Yao, J.P.A. Dewald / NeuroImage 25 (2005) 369–382 373

Statistical analysis method Fig. 3). It is worth to note that when applying MDP2 to simulated
data set III, there is a decreasing trend in the location error when
The three evaluation indices (i.e., ED, PUS, and PFS) were the distance between the two real dipoles increases. This trend is
subjected to separate multivariate analysis of variance (MAN- not obvious for any of the other methods. This result suggested that
OVA). The two dependent variables of the MANOVA were the the a priori knowledge of the number of dipoles could improve the
mean and standard deviation of each tested index, and the three inverse results when using dipole source models. Unfortunately,
factors were the inverse method, the type of simulated EEG data even when the correct number of dipoles is implemented, the
set, and the radius of the dipole area (for data set I/II) or the MDP2 method still produces posterior and medial shifts, which
distance between the two dipoles (for data set III). Up to 2-way result in substantial location errors.
interactions between the factors were considered when performing The results of statistical analysis using a three-way MANOVA
the analysis. The post hoc analysis was implemented using the and Bonferroni post hoc tests on the mean and standard deviation
Bonferroni test at the P = 0.05 level of significance. of location errors are shown in Table 2a. In this table, values in the
upper and lower triangle are P values on the mean and standard
deviation of different indices, respectively. Results in Table 2a
Results show that location estimated by LRT1 method is significantly
more accurate than any other methods tested in this paper (see
Source localization results obtained on simulated EEG data statistical results on ED in the top triangle of Table 2a). Further-
more, the variability (estimated by standard deviation) of the LRT1
Example results obtained using each of the methods on method on estimating the location of source is either significantly
simulated data sets I/II/III (with a radius that equals 8 mm for less than or similar to other methods (see statistical results on the
data sets I and II, and with a distance between the two dipoles that standard deviations in the bottom triangle of Table 2a). The
equals 18 mm for data set III) are shown in Figs. 2A–C, location error obtained by using LRT15 is significantly smaller
respectively. Results are shown in a brain coordinate system with than results obtained by using any other methods except LRT1 and
x- and y-axis representing anterior (negative x)/posterior (positive MDP1.
x) and left (positive y)/right (negative y) in mm. In this figure, the
first and second rows showed the results of minimum Lp norm ( P = The percentage of undetected source number
1, 1.5, 2) and LRT Lp norm ( P = 1, 1.5, 2), respectively. Color bars The percentages of undetected source number (PUS) of each
on the right side of each subplot show the averaged current strength source localization method using simulated data sets I/II/III are
over the 10 time points on the cortex in AAmm. On the last row, the shown in Fig. 4. Since dipole methods are not able to detect the
purple dots on the first two subplots are the mean of the estimated area of the source, the PUSs for dipole methods are close to 100%.
dipole location over the 10 time points obtained by using MDP1 When using simulated data sets I and II, PUSs of all CDR methods
and MDP2 methods. The horizontal, vertical, diagonal lines have a tendency to increase when the size of dipole area increases.
represent the standard deviation of estimated dipole position on This tendency is slower for the LRT1 method than that for any of
x, y, and z directions. Finally, the purple dots in the last subplot the other tested CDR methods. It is worth noting that when using
represent the real source dipole positions. Quantitative evaluation simulated data set II to evaluate the CDR methods, there is an
results and statistical analysis results were performed and will be obvious drop of PUS when the radiuses of the two source dipole-
discussed next. areas increase from 5 mm to 6 mm. When referring to results
obtained with simulated data set II with the radiuses of the two
Location accuracy dipole areas equal to 5 mm in Fig. 2d, one finds that all the CDR
The location error distances (EDs) of each inverse method methods could not detect the laterally located dipole area; instead,
using simulated data sets I/II/III are shown in Fig. 3. In this figure, they detected a single enlarged and merged dipole area located
the gray bars on each of the data points represent the standard error closer to the medial dipole area (see Fig. 2d). When increasing the
of corresponding results. Results clearly show that the LRT1 radiuses from 5 mm to 6 mm, all CDR methods show an un-
method provides the best estimation of source location for all three separated bimodal area, which is related to the drop of PUS
simulated EEG data sets. When applying CDR methods on observed in Fig. 4b. This result shows that a 5 mm source grid does
simulated data sets I and II, the EDs of LRT1 have general not allow for the detection of two separated sources each with a 5
decreasing trends when the sizes of dipole-areas increase. How- mm radius and with 29 mm between their centers.
ever, such a decreasing trend is not so obvious for the other The simulated data set III consists of two dipoles. For this type
methods on both data sets I and II (see Figs. 3a and b). Results of of EEG data, all CDR methods are able to detect the two dipoles
EDs obtained by applying LRT with Lp ( P = 1, 1.5, 2) norm when the distance between the 2 source dipoles is less than 12 mm.
methods on simulated data set III fluctuate when the distance However, most of CDR methods except LRT1 are unable to detect
between the two source dipoles is about 12–16 mm (see Fig. 3c). one of the two dipoles when the distance between the 2 source
The reason for this fluctuation is believed to be related to dipoles is larger than 12 mm. In all the circumstances, LRT1 has
constraints imposed on source location by the 5 mm grid on the the lowest PUS from all CDR methods.
cortex (see Materials and methods). The results of statistical analysis using three-way MANOVA
Although the moving dipole methods (MDP1 and MPD2) show and Bonferroni post hoc tests on the mean and standard deviation
relatively low EDs, there is an obvious posterior and medial shift of the percentages of undetected source number (PUSs) are shown
shown on the results of all the three simulated EEG data sets, while in Table 2b. Results show that LRT1 has a significantly smaller and
such a shift is not found in the results obtained by using CRD less variable PUS than all the other methods tested in this paper
methods (see Fig. 2). Even worse, locations estimated by the (except that the mean and standard deviation on PUSs for LRT1
MDP2 method over 10 time points show a large variance (see and LRT15 are similar).
374 J. Yao, J.P.A. Dewald / NeuroImage 25 (2005) 369–382

The percentage of falsely-detected source number are shown in Fig. 5. Results show that when using simulated data
The percentages of falsely-detected source number (PFSs) of sets I and II, PFSs of LRT1 and LRT15 methods have a tendency to
each source localization method using simulated data sets I/II/III decrease when the size of dipole area increases, while this trend is

Fig. 2. Results obtained using each of the investigated inverse methods for simulated EEG data with (A) single dipole area with a radius equal to 8 mm, two areas
of dipoles with a distance between two centers equal to 29 mm and radiuses equal to (D) 5 mm and (B) 8 mm, and (C) two single source dipoles with an inter-
dipole distance of 18 mm. The order of the inverse methods is corresponding to the methods 1–8 in Fig. 1. Results were shown in brain coordinates with x- and y-
axis representing anterior (negative x)/posterior (positive x) and left (positive y)/right (negative y) in mm. Color bars on the right side of subplots in the first two
rows show the averaged current strength over the 10 time points on the cortex in AAmm. In the last row, the purple dots on the first two subplots are the mean of
the estimated dipole locations over the 10 time points obtained by using MDP1 and MDP2 methods. The horizontal, vertical, and diagonal lines represent the
standard deviation of estimated dipole position on x, y, and z directions. Finally, the purple dots in the last subplot represent the source dipole positions.
J. Yao, J.P.A. Dewald / NeuroImage 25 (2005) 369–382 375

Fig. 2 (continued).

not apparent for the other CDR methods. PFSs of MDP1 for MDP2 has a significant larger standard deviation compared to
simulated data sets I/II/III are always equal to 100%, which shows other methods (see Table 2c).
that this method is not able to detect the source location within the The results of statistical analysis using three-way MANOVA
source area (caused by the posterior and medial shift). The PFSs of and Bonferroni post hoc tests on the mean and standard deviation
MDP2 decrease with the increase of the radius of source area using of the percentages of PFS are shown in Table 2c. Results showed
simulated data set I, which clearly suggests that adding a second that LRT1 has a significantly smaller PFS than any of the other
dipole is beneficial in reducing the false detection ratio. However, methods.
376 J. Yao, J.P.A. Dewald / NeuroImage 25 (2005) 369–382

the sources of SEPs are located on the primary sensory cortex (S1
area). Results obtained by using CDR methods have a similar dhot
spotT located on S1 area, while multiple other active sites besides S1
area also show up. Among all the CDR methods, results of methods
L1, L15, and LRT1 are more focused on S1 area; furthermore,
methods L15 and LRT1 show relatively stronger activity on S1 area.
Results of ERPs related to self-initiated isometric shoulder
abduction during a time window from 110 to 90 ms before the
onset of torque are shown in Fig. 7. The majority of activities prior
to shoulder abduction obtained by all the CDR methods are located
on the contralateral hemisphere. Results by MDP1 and MDP2 are
located either at prefrontal lobe or parietal lobe. Activities on
prefrontal and parietal lobes are also found on the results obtained
by CDR methods. However, one observes that the result obtained
by LRT1 is the most focused on the medial pre-central gyrus with
very limited activity anywhere else. The L1 CDR method
performed only slightly worse than LRT1. All other CDR methods
show activity in many regions other than M1. The results obtained
with LRT1 are consistent with those reported in the previous

Table 2
Results of Bonferroni post hoc tests on the mean and standard deviation of
the ED, PUS, and PFS indices

Fig. 3. The error distance for the three simulated EEG data sets using the
eight inverse methods. The gray error bars on each of the data points
represent the standard error.

Source localization results on real EEG data

The results of different inverse methods on SEPs resulting from


electrical stimulation of the subject’s wrist during the time window
from 20 to 25 ms after the stimulus are represented in Fig. 6. As
shown in Fig. 6, the majority of activities obtained by all the
methods stay on the contralateral hemisphere. Inverse results on
SEPs from shoulder, elbow, and wrist shift from medial to lateral
along the central sulcus. (SEP results of simulating elbow and
shoulder are not represented here.) Qualitatively, one can observe
In this table, values in the upper (white cells) and lower triangle (gray cells)
that the inverse results obtained by using method MDP1 or MDP2 are P values of the mean and standard deviation, respectively. The table
have a large variance even during a short time window (5 ms). And should be read from row to column, e.g., the element in first row and third
the mean locations obtained by MDP1 and MDP2 are located at pre- column of (c) shows that the mean of PFS for the L1 method is significantly
central gyrus (i.e., motor cortex). These results are not consistent larger (+) than the mean for the L2 method, where +/P b 0.05, ++/P b
with results published in the previous literature, which suggests that 0.01, +++/P b 0.001, ++++/P b 0.0001.
J. Yao, J.P.A. Dewald / NeuroImage 25 (2005) 369–382 377

Although the evaluation on the source localization ability of two or


three different methods was already reported previously, this is the
first paper that compared these many inverse methods using both
simulated and experimental data sets. Our evaluation results
provide useful information for choosing the appropriate cortical
source localization techniques for scientific and clinical purposes.

Evaluation methods

In order to evaluate the localization ability of the different


source localization methods, we designed three types of simu-
lated EEG data sets. Results obtained with each of the simulated
EEG data sets represented various facets of the accuracy of the
inverse methods. Simulated data sets I and II allowed for the

Fig. 4. The percentage of undetected source number for the three simulated
EEG data sets using the eight inverse methods. The gray error bars on each
of the data points represent the standard error.

literature and demonstrated that M1 is the primary source of


activity related to arm activity.
In general, inverse results obtained by applying LRT1 to
SEPs and ERPs are qualitatively better than other methods,
because it provides results that mirror those obtained with direct
cortical recordings or with other noninvasive neuroimaging
modalities.

Discussions and Conclusions

In order to evaluate different cortical source localization Fig. 5. The percentage of false-detected source number for the three
methods, we applied 8 different inverse methods available in simulated EEG data sets using the eight inverse methods. The gray bars on
CURRY software to both simulated and experimental EEG data. each of the data points represent the standard error.
378 J. Yao, J.P.A. Dewald / NeuroImage 25 (2005) 369–382

Fig. 6. Inverse calculation results for SEPs resulting from electrical stimulation of the medial nerve at wrist in a subject. The figure legends are the same as
those in Fig. 2. The green line in each of the subplots is the central sulcus.

evaluation of source localization ability; and data sets II and III experimental EEG data sets (i.e., SEPs and ERPs) were used to
allowed for the evaluation of the spatial resolution of the various qualitatively evaluate results obtained with each of the inverse
techniques. methods. Our assessments are based on our current knowledge of
Besides simulated EEG data, we also evaluated the different regions of cortical sensory motor activity associated with sensory
inverse methods using experimental EEG data. Two typical evoked or motor-related potentials.

Fig. 7. Inverse calculation results for ERPs related to isometric shoulder abduction torque generated by a subject. The figure legends are the same as those in
Fig. 2. The green line in each of the subplots is the central sulcus.
J. Yao, J.P.A. Dewald / NeuroImage 25 (2005) 369–382 379

In case of applying inverse methods to simulated EEG data, of dipole strength was included. The square-wave-shaped sources
where the real source locations are known, three different used here are sub-optimal for LRT Lp norm methods, which are
indices, i.e., the error distance (ED), percentage of undetected designed on the assumption that the source activities are distributed
source number (PUS), and percentage of falsely-detected source smoothly on the source layer. We postulate that using simulated
number (PFS), were used to quantify the performance of the EEG data with decays will further improve the performance of
different inverse methods. A method with lower ED, PUS, and LRT Lp norm methods.
PFS is able to detect the source location more accurately. Based When evaluating the CDR methods, a threshold, which was set
on the definition of ED, there is a positive correlation between at a 98% confidence level, was used to define a source. It is
ED and PUS/PFS, i.e., a larger PUS or PFS will cause an increase obvious that the choice of threshold also affects our evaluation
in ED. The correlation coefficients between ED and PUS/PFS are results. Since currents estimated by different CDR methods have
shown in Table 3. Because of the highly significant positive different distributions, we choose a data-specific threshold for each
correlation between ED and PUS/PFS, one can only refer to ED of the inverse results (i.e., we first fitted the estimated cortical
for the evaluation of the source localization ability of an inverse currents using a Weibull distribution; then we set the threshold at
method. the 98% confidence level). Even with such a threshold, the
residuals of the fitting procedure are slightly different depending
Factors that may influence our evaluation methods on results obtained by different localization methods. Therefore,
there possibly is a small bias for ED, PUS, and PFS depending on
Several factors that may influence the performance of inverse the choice of source localization method.
methods were not studied in this paper. First of all, previous reports Using a data-specific threshold at 98% confidence level, all
showed that results of inverse calculation are sensitive to the CDR methods tend to provide a relatively large PFS; however,
conductivity of head (Benar and Gotman, 2002; Gencer and Acar, different methods have a different distribution of the falsely-
2004; van Burik and Peters, 2000; van den Broek et al., 1998; Wen detected sources. For example, when using LRT1 method, the
and Li, 2001). Among conductivities of brain, skull, and scalp, the falsely-detected sources are distributed around the real source;
inverse result is most sensitive to the conductivity of skull (Ferree while when using other CDR methods, more falsely-detected
et al., 2000). However, the real conductivity of skull in vivo is still sources are distributed around the edge of the cortex (see Fig. 8,
unknown. Direct measurements showed a large variance on this where the purple crosses represent the detected source positions).
value ranging from 15 mS/m to 80 mS/m (Akhtari et al., 2002; By increasing the threshold, we can effectively increase the PUS
Ferree et al., 2000; Hoekema et al., 2003). The conductivity values and decrease the PFS of CRD methods. Since a positive correlation
used in this paper were listed in Table 1 and corresponded to a existed both between ED and PUS as well as between ED and PFS,
skull-scalp-ratio equal to 1:15. Inverse results were also obtained the impact of an increased threshold on ED is not straightforward.
using skull-scalp-ratios equal to 1:26 and 1:80. In all of these cases, Other factors, such as the number of electrodes, the depth of the
LRT1 provides the overall best results. source, and the SNR of signals, also have impact on the
Secondly, the regulation parameter, k, is crucial to the performance of the various inverse methods. The impact of these
performance of inverse methods. As a general rule, the degree parameters were not included as part of the current investigation.
of regularization (k) should increase with the level of noise in the
data. Based on this rule, k is often determined using v 2 and L- Accuracy of the different source localization methods
curve methods. Unfortunately, these two methods are not
available for the minimum Lp norm and LRT Lp norm methods According to the results obtained using 3 different simulated
in CURRY. In this paper, a fixed value (k = 1) was used for all EEG data sets, the LRT L1 norm method showed the overall
the methods. In order to limit the potential influence caused by highest accuracy. We further summarized the comparison on the
using a non-optimized regulation parameter, we controlled the mean and the standard deviation (inside the parentheses) of the
signal to noise ratio (SNR) of all the EEG data, including both the three evaluation indices between LRT1 to other inverse methods in
simulated and experimental data, to a very consistent level, Table 4. (Note, only significant results were represented.) As
ranging from 9 to 10. Regardless, evaluation of the different shown in Table 4, the LRT1 method has either significant better or
source localization methods by using an optimized regulation similar performance when compared to the other investigated
parameter remains desirable. methods for all of the three indices, which showed an overall
The shape of simulated EEG data may be another factor that superior ability to detect the location of the source. When applying
will affect the evaluation results. All the three simulated EEG data the various methods to experimental EEG data, including both
sets are composed of equal strength dipoles (1 AAmm). No decay SEPs and ERPs, the LRT1 method also provides the best estimates.
These findings further confirm that among all the tested methods,
LRT1 has the most accurate and least variable source localization
Table 3
abilities.
Pearson product moment correlations between the three evaluation indices
and P values obtained using ANOVA for linear regression Our results further indicate that a weighted minimum norm
method, such as LRT-Lp, has better source localization ability than
Simulated EED Simulated EED Simulated EED
the regular minimum norm methods. These results are in
data I data II data III
accordance with those reported in previous literature (Baillet and
PUS PFS PUS PFS PUS PFS Garnero, 1997; Babiloni et al., 2000b, 2004). Furthermore, we also
ED 0.704 0.629 0.864 0.718 0.362 0.245 demonstrate that CDR methods provide physiologically meaningful
(b0.0001) (b0.0001) (b0.0001) (0.0386) (0.02) (0.350) results while MDP1 and MDP2 solutions failed in many situations,
In this table, P values and R values are shown inside and outside the which is also in agreement with previous reported results (He et al.,
parentheses, correspondingly. 2002a; Pascual-Marqui et al., 1994). Within the same CDR
380 J. Yao, J.P.A. Dewald / NeuroImage 25 (2005) 369–382

Fig. 8. Inverse calculation results for simulated EEG data set II with radius of the two dipole areas equal to 8 mm. Figure legends were the same as those in
Fig. 2. The purple dots represent the estimated and real source dipole positions.

regulation method (minimum norm or weighted minimum norm) a the method is defined as the minimum radius of dipole area that a
lower Lp value will improve the localization results. method can distinguish between the two areas. Simulated EEG data
set III consists of two moving dipoles with the distance between
Spatial resolution of different methods them increasing from 4 mm to 20 mm. We defined the spatial
resolution of an inverse method applied to this type of data as the
The purpose of applying inverse methods is to improve the minimum distance between two moving dipoles that can be
spatial resolution of EEG measurement. However, to date, no distinguished.
quantitative examination has been performed to provide the spatial Results on simulated EEG data II showed that when the radiuses
resolution of an inverse method. For this reason, we designed of the two dipole areas are equal to 5 mm, none of the CDR methods
simulated EEG data sets II and III to quantify the spatial resolution could separate the two areas (i.e., only one mode is detected). When
of different inverse methods for 163-channel EEG signals with a increasing the radiuses from 5 mm to 6 mm, L15, L1, LRT15, and
SNR around 9–10. Simulated EEG data set II has two dipole-areas LRT1 methods showed an un-separated area, however, with a
with increasing radiuses (ranging from 5 mm to 10 mm) and a bimodal distribution. Furthermore, two separated areas can be found
fixed distance (29 mm) between their centers. When applying an when using a higher threshold. These results suggest that the spatial
inverse method to simulated EEG data II, the spatial resolution of resolution for these methods is around 6 mm. Further increasing the
radiuses of the two dipole areas did not cause the merge of the two
Table 4 dipole areas. This is a very interesting result. Considering that by
Comparison of the three evaluation indices between LRT1 and the other increasing the radiuses of the two dipole areas while keeping the
methods distance between their centers the same, the two source regions were
Methods LRT1 getting closer, which we thought would require a better spatial
resolution to separate them. However, this is not correct. A source
ED PUS PFS
image with a smaller radius has higher frequency components. In
L1 ++++ (+++) ++++ (++++) ++ (+++) the extreme case, the frequency component of a source image with
L15 ++++ ++++ (+++) ++++ only a single dipole is the highest. A source containing higher
L2 ++++ ++++ (++++) ++++
frequency components requires a higher spatial resolution to detect.
LRT15 ++++ ++
LRT2 ++++ (+) ++++ (++++) ++++
Therefore, our results showed a better ability to separate two areas
MDP1 ++++ ++++ ++++ when the radiuses of the two source dipole-areas are increasing.
MDP2 ++++ (++++) ++++ ++++ (++++) The effect of the frequency component of the image would also
explain the results obtained using simulated EEG data set III which
The table should be read from row (method) to column (index), e.g., the
element in first row and second column shows that the mean of ED for L1
showed that when increasing the distance between the two dipoles,
method is significantly larger than the mean of ED for LRT1 method, where no obvious change in performance of the CDR methods was
+/P b 0.05, ++/P b 0.01, +++/P b 0.001, ++++/ P b observed ( P values of post hoc analysis on the effect of radius are
0.0001. The results with and without parentheses represent the P value of larger than 0.05). Obviously, changing the distance between the
the standard deviation and mean, respectively. two dipoles does not change the frequency component of source
J. Yao, J.P.A. Dewald / NeuroImage 25 (2005) 369–382 381

image enough to improve the performance of the CDR source Beer, R.F., Johnston, E.R., Eisenberg, E., 1995. Vector Mechanics for
localization methods. Engineers: Statistics. McGraw-Hill, New York.
Another aspect that may have influenced our results is the use Benar, C.G., Gotman, J., 2002. Modeling of post-surgical brain and skull
defects in the EEG inverse problem with the boundary element method.
of the 5 mm grid of source layer to constrain the source locations.
Clin. Neurophysiol. 113, 48 – 56.
We believe that by using a finer grid the spatial resolution for L15,
Cheyne, D., Kristeva, R., Deecke, L., 1991. Homuncular organization of
L1, LRT15, and LRT1 methods would further improve the human motor cortex as indicated by neuromagnetic recordings. Neuro-
performance of these CDR methods. sci. Lett. 122, 17 – 20.
Cincotti, F., Babiloni, C., Miniussi, C., Carducci, F., Moretti, D., Salinari,
Practical implications for the application of the various source S., Pascual-Marqui, R., Rossini, P.M., Babiloni, F., 2004. EEG
localization methods deblurring techniques in a clinical context. Methods Inf. Med. 43,
114 – 117.
Dipole source models are more widely used in the previous Dowman, R., Schell, S., 1999. Innocuous-related sural nerve-evoked and
literature to estimate centers of cortical activity. We believe that finger-evoked potentials generated in the primary somatosensory and
they are only suitable for the situation where a single source is supplementary motor cortices. Clin. Neurophysiol. 110, 2104 – 2116.
Druschky, K., Kaltenhauser, M., Hummel, C., Druschky, A., Huk, W.,
presented. In order to model multiple sources, multiple dipole
Neundorfer, B., Stefan, H., 2003. Somatosensory evoked magnetic
methods were developed. However, our results showed that even fields following passive movement compared with tactile stimulation of
using the correct number of dipoles, multiple dipole methods could the index finger. Exp. Brain Res. 148, 186 – 195.
not determine the accurate location of all the sources. Emerson, R., Pedley, T., 1984. Generator sources of median somatosensory
CDR methods have the ability to determine the current evoked potentials. J. Clin. Neurophysiol. 1, 203 – 218.
distribution of brain activity. Among all the CDR methods Ferree, T.C., Eriksen, K.J., Tucker, D.M., 2000. Regional head tissue
investigated in this paper, the LRT1 method showed the highest conductivity estimation for improved EEG analysis. IEEE Trans.
accuracy and, therefore, is the most promising method for source Biomed. Eng. 47, 1584 – 1592.
localization estimation especially when different cortical regions Ferrier, D., 1875. Experiments on the brain of monkeys. Proc. R. Soc.
are expected to be simultaneously active during a particular London, B: Biol. Sci. 24, 409 – 432.
Fuchs, M., Drenckhahn, R., Wischmann, H.-A., Wager, M., 1998. An
experimental paradigm.
improved boundary element method for realistic volume-conductor
modeling. IEEE Trans. Biol. Med. Eng. 45, 980 – 997.
Gencer, N.G., Acar, C.E., 2004. Sensitivity of EEG and MEG measure-
Acknowledgments ments to tissue conductivity. Phys. Med. Biol. 49, 701 – 717.
Haan, H., Streb, J., Bien, S., Rosler, F., 2000. Individual cortical current
This work was supported by a National Institutes of Health density reconstructions of the semantic N400 effect: using a generalized
Grant 1R03HD39804-01A1 and by a donation from the RIC minimum norm model with different constraints (L1 and L2 norm).
Women’s Board. The authors want to acknowledge the construc- Hum. Brain Mapp. 11, 178 – 192.
tive suggestions made by Drs. Leo Towle, Wim van Drongelen, Hamalainen, M.S., Ilmoniemi, R.J., 1994. Interpreting magnetic fields of the
Aryre Nehorai, and the reviews of NeuroImage during the brain: minimum norm estimates. Med. Biol. Eng. Comput. 32, 35 – 42.
Hari, R., Forss, N., 1999. Magnetoencephalography in the study of human
preparation and revision of the manuscript.
somatosensory cortical processing. Philos. Trans. R. Soc. London, B:
Biol. Sci. 354, 1145 – 1154.
Hashimoto, I., 2000. High-frequency oscillations of somatosensory evoked
References potentials and fields. J. Clin. Neurophysiol. 17, 309 – 320.
He, B., Yao, D., Lian, J., Wu, D., 2002a. An equivalent current source
Akhtari, M., Bryant, H.C., Mamelak, A.N., Flynn, E.R., Heller, L., Shih, model and Laplacian weighted minimum norm current estimates of
J.J., Mandelkern, M., Matlachov, A., Ranken, D.M., Best, E.D., brain electrical activity. IEEE Trans. Biomed. Eng. 49, 277 – 288.
DiMauro, M.A., Lee, R.R., Sutherling, W.W., 2002. Conductivities of He, B., Zhang, X., Lian, J., Sasaki, H., Wu, D., Towle, V.L., 2002b.
three-layer live human skull. Brain Topogr. 14, 151 – 167. Boundary element method-based cortical potential imaging of somato-
Asanuma, H., Rosen, I., 1972. Topographical organization of cortical sensory evoked potentials using subjects’ magnetic resonance images.
efferent zones projecting to distal forelimb muscles in the monkey. Exp. NeuroImage 16, 564 – 576.
Brain Res. 14, 243 – 256. Hines, M., 1940. Movements elicited from precentral gyrus of adult
Babiloni, F., Babiloni, C., Locche, L., Cincotti, F., Rossini, P., Carducci, F., chimpanzees by stimulation with sine wave currents. J. Neurophysiol.
2000a. High-resolution electro-encephalogram: source estimates of 3, 442 – 466.
Laplacian-transformed somatosensory-evoked potentials using a real- Hoekema, R., Wieneke, G.H., Leijten, F.S., van Veelen, C.W., van Rijen,
istic subject head model constructed from magnetic resonance images. P.C., Huiskamp, G.J., Ansems, J., van Huffelen, A.C., 2003. Measure-
Med. Biol. Eng. Comput. 38 (5), 512 – 519. ment of the conductivity of skull, temporarily removed during epilepsy
Babiloni, F., Babiloni, C., Locche, L., Cincotti, F., Rossini, P.M., Carducci, surgery. Brain Topogr. 16, 29 – 38.
F., 2000b. High-resolution electro-encephalogram: source estimates of Ilmoniemi, R.J., 1993. Models of source currents in the brain. Brain Topogr.
Laplacian-transformed somatosensory-evoked potentials using a real- 5, 331 – 336.
istic subject head model constructed from magnetic resonance images. Komssi, S., Huttunen, J., Aronen, H.J., Ilmoniemi, R.J., 2004. EEG
Med. Biol. Eng. Comput. 38, 512 – 519. minimum-norm estimation compared with MEG dipole fitting in the
Babiloni, F., Babiloni, C., Carducci, F., Romani, G.L., Rossini, P.M., localization of somatosensory sources at S1. Clin. Neurophysiol. 115,
Angelone, L.M., Cincotti, F., 2004. Multimodal integration of EEG and 534 – 542.
MEG data: a simulation study with variable signal-to-noise ratio and Nenonen, J.T., Hamalainen, M.S., Ilmoniemi, R.J., 1994. Minimum-norm
number of sensors. Hum. Brain Mapp. 22, 52 – 62. estimation in a boundary-element torso model. Med. Biol. Eng.
Baillet, S., Garnero, L., 1997. A Bayesian approach to introducing Comput. 32, 43 – 48.
anatomo-functional priors in the EEG/MEG inverse problem. IEEE NeuroScan, L., 1999. Curry Reference Guide. pp. 101 – 109.
Trans. Biomed. Eng. 44, 374 – 385. Nunez, P., 1981. Quantitative states of neocortex. In: Nunez, P. (Ed.),
382 J. Yao, J.P.A. Dewald / NeuroImage 25 (2005) 369–382

Neocortical Dynamics and Human EEG Rhythms. Oxford University Stenbacka, L., Vanni, S., Uutela, K., Hari, R., 2002. Comparison of mini-
Press, Inc., New York, pp. 3 – 67. mum current estimate and dipole modeling in the analysis of simulated
Pascual-Marqui, R.D., Biscay-Lirio, R., 1993. Spatial resolution of neuro- activity in the human visual cortices. NeuroImage 16, 936 – 943.
nal generators based on EEG and MEG measurements. Int. J. Neurosci. Towle, V.L., Khorasani, L., Uftring, S., Pelizzari, C., Erickson, R.K., Spire,
68, 93 – 105. J.P., Hoffmann, K., Chu, D., Scherg, M., 2003. Noninvasive identifica-
Pascual-Marqui, R.D., Michel, C.M., Lehmann, D., 1994. Low resolution tion of human central sulcus: a comparison of gyral morphology, func-
electromagnetic tomography: a new method for localizing electrical tional MRI dipole localization, and direct cortical mapping. NeuroImage.
activity in the brain. Int. J. Psychophysiol. 18, 49 – 65. Valeriani, M., Le Pera, D., Niddam, D., Arendt-Nielsen, L., Chen, A., 2000.
Penfield, W., Rasmussen, T., 1950. The cerebral cortex of man. A Clinical Dipolar source modeling of somatosensory evoked potentials to painful
Study of Localization of Function. Macmillan, New York. and nonpainful median nerve stimulation. Muscle Nerve 23, 1194 – 1203.
Phillips, C., Rugg, M.D., Fristont, K.J., 2002. Systematic regularization of van Burik, M.J., Peters, M.J., 2000. Estimation of the electric conductivity
linear inverse solutions of the EEG source localization problem. from scalp measurements: feasibility and application to source local-
NeuroImage 17, 287 – 301. ization. Clin. Neurophysiol. 111, 1514 – 1521.
Rao, S.M., Binder, J.R., Hammeke, T.A., Bandettini, P.A., Bobholz, J.A., van den Broek, S.P., Reinders, F., Donderwinkel, M., Peters, M.J., 1998.
Frost, J.A., Myklebust, B.M., Jacobson, R.D., Hyde, J.S., 1995. Volume conduction effects in EEG and MEG. Electroencephalogr. Clin.
Somatotopic mapping of the human primary motor cortex with Neurophysiol. 106, 522 – 534.
functional magnetic resonance imaging. Neurology 45, 919 – 924. Waberski, T.D., Buchner, H., Perkuhn, M., Gobbele, R., Wagner, M.,
Recuero, E., 1999. The use of somatosensory evoked potentials for cortical Kucker, W., Silny, J., 1999. N30 and the effect of explorative finger
localization during surgical operations. Rev. Neurol. 28, 588 – 590. movements: a model of contribution of the motor cortex to early
Restuccia, D., Valeriani, M., Insola, A., Lo Monaco, M., Grassi, E., Barba, somatosensory potentials. Clin. Neurophysiol. 110, 1589 – 1600.
C., Le Pera, D., Mauguiere, F., 2002. Modality-related scalp responses Waberski, T.D., Gobbele, R., Herrendorf, G., Steinhoff, B.J., Kolle, R.,
after electrical stimulation of cutaneous and muscular upper limb Fuchs, M., Paulus, W., Buchner, H., 2000. Source reconstruction of
afferents in humans. Muscle Nerve 26, 44 – 54. mesial-temporal epileptiform activity: comparison of inverse techni-
Ricamato, A., Dhaher, Y., Dewald, J., 2003. Estimation of active cortical ques. Epilepsia 41, 1574 – 1583.
current source regions using a vector representation scanning approach. Wagner, M., Kohler, T., Fuchs, M., Kastner, J., 2000. An extended
J. Clin. Neurophysiol. 20, 326 – 344. source model for current density reconstructions. In: Nenonen, J.,
Salenius, S., Portin, K., Kajola, M., Salmelin, R., Hari, R., 1997. Cortical Ilmoniemi, R.J., Katila, T. (Eds.), Proc. 12th Int. Conf. Biomagnetism,
control of human motoneuron firing during isometric contraction. J. Biomag2000. Helsinki University of Technology, Espoo, Finland,
Neurophysiol. 77, 3401 – 3405. pp. 749 – 752.
Sanes, J.N., Donoghue, J.P., Thangaraj, V., Edelman, R.R., Warach, S., Wang, J.Z., Williamson, S.J., Kaufman, L., 1992. Magnetic source images
1995. Shared neural substrates controlling hand movements in human determined by a lead-field analysis: the unique minimum-norm least-
motor cortex. Science 268, 1775 – 1777. squares estimation. IEEE Trans. Biomed. Eng. 39, 665 – 675.
Schaefer, M., Muhlnickel, W., Grusser, S., Flor, H., 2002. Reproducibility Wang, J.Z., Kaufman, L., Wiliamson, S.J., 1993. Imaging regional changes
and stability of neuroelectric source imaging in primary somatosensory in the spontaneous activity of the brain: an extension of the minimum-
cortex. Brain Topogr. 14, 179 – 189. norm least-squares estimate. Electroencephalogr. Clin. Neurophysiol.
Scherg, M., 1990. Fundamentals of dipole source potential analysis. In: 86, 36 – 50.
Gandori, F., Hoke, M., Romani, G.L. (Eds.), Auditory Evoked Magnetic Weibull, W., 1939. A statistical theory of the strength of materials.
Fields and Potentials: Advances in Audiology, vol. 6. Karger, Basel, Ingeniors Vetenskaps Akademiens Handlingar, R. Swed. Inst. Eng.
pp. 40 – 70. Res., vol. 153. Stockholm, Sweden.
Scherg, M., Buchner, H., 1993. Somatosensory evoked potentials and Wen, P., Li, Y., 2001. Comparison study of different head model structures
magnetic fields: separation of multiple source activities. Physiol. Meas. with homogeneous/inhomogeneous conductivity. Australas. Phys. Eng.
14 (Suppl. 4A), A35 – A39. Sci. Med. 24, 31 – 36.
Schieber, M.H., 2001. Constraints on somatotopic organization in the Williamson, S.J., Kaufman, L., 1981. Biomagnetism. J. Magn. Magn. Mat.
primary motor cortex. J. Neurophysiol. 86, 2125 – 2143. 22, 129 – 201.
Schieber, M.H., Hibbard, L.S., 1993. How somatotopic is the motor cortex Woolsey, C.N., 1958. Organization of somatic sensory and motor areas of
hand area? Science 261, 489 – 492. the cerebral cortex. In: Warlow, H.F., Woolsey, C.N. (Eds.), The
Sherrington, C.S., 1947. The Integrative Action of the Nervous System Biological and Biochemical Bases of Behavior. University of Wisconsin
(1906), 2nd ed. Cambridge University Press (reprinted 1973). Press, Madison, pp. 63 – 81.
Sonoo, M., 2000. Anatomic origin and clinical application of the Yao, J., Dewald, J.P.A., 2003. Evaluation of different cortical potential
widespread N18 potential in median nerve somatosensory evoked imaging methods using simulated EEG data. IEEE EMBS, Proc. 25th
potentials. J. Clin. Neurophysiol. 17, 258 – 268. Ann. Int. Conf. IEEE, Cancun, Mexico, pp. 2128 – 2131.

Вам также может понравиться