ArticlePDF Available

New thin-film surface electrode array enables brain mapping with high spatial acuity in rodents

Authors:

Abstract and Figures

In neuroscience, single-shank penetrating multi-electrode arrays are standard for sequentially sampling several cortical sites with high spatial and temporal resolution, with the disadvantage of neuronal damage. Non-penetrating surface grids used in electrocorticography (ECoG) permit simultaneous recording of multiple cortical sites, with limited spatial resolution, due to distance to neuronal tissue, large contact size and high impedances. Here we compared new thin-film parylene C ECoG grids, covering the guinea pig primary auditory cortex, with simultaneous recordings from penetrating electrode array (PEAs), inserted through openings in the grid material. ECoG grid local field potentials (LFP) showed higher response thresholds and amplitudes compared to PEAs. They enabled, however, fast and reliable tonotopic mapping of the auditory cortex (place-frequency slope: 0.7 mm/octave), with tuning widths similar to PEAs. The ECoG signal correlated best with supragranular layers, exponentially decreasing with cortical depth. The grids also enabled recording of multi-unit activity (MUA), yielding several advantages over LFP recordings, including sharper frequency tunings. ECoG first spike latency showed highest similarity to superficial PEA contacts and MUA traces maximally correlated with PEA recordings from the granular layer. These results confirm high quality of the ECoG grid recordings and the possibility to collect LFP and MUA simultaneously.
The newly developed ECoG grid enabled recordings of evoked MUA similar to conventional penetrating multi-electrode arrays. (A) Examples of filtered (300-3000 Hz) traces with detected spikes (red circles) above detection threshold (dashed line) for both ECoG and penetrating multi-electrode array (PEA). Stimulus onset is indicated (red arrow head). (B) Representative examples of averaged LFP and raster-plots of MUA in response to a broad-band noise stimulus of 80 dB SPL rms. The raster-plots indicate the time points of every detected spike in each of the 30 repetitions. Stimulus onset is indicated (vertical line). (C) The peristimulus time histograms (psth) show representative responses to a 100 ms broad-band stimulus, for the ECoG grid and the penetrating electrode (PEA), respectively. Usually a strong onset response and a weak offset response were discernable. (D) Similarity between first spike latency (FSL) for ECoG grid and penetrating multi-electrode array contacts decreases with cortical depth (depth3 = infragranular layer). Given are median FSLs at the best frequency threshold level (i.e. 10% of max MUA, comparable to the CF threshold) for ECoG grid recording sites close to the insertion point and for different depths at the PEA (depth 1: 0-600 µm, depth 2: 750-1200 µm, depth 3: 1350-1800 µm). Box-plots with min and max values (whisker) and mean (cross). The FSL at the surface was significantly shorter than the one recorded in deep layers of the AC; dependent t-tests with Bonferroni correction *p < 0.05. (E) Correlation strength (r 2 ) between grid and penetrating electrode recordings for MUA is highest for granular layers (contact #7). Given are mean (dot) and SEM (whisker). After an initial decline in correlation strength, the maximal correlation was reached at the 7 th contact and sharply declined with increasing cortical depth (contacts #14-16, indicated in light grey, were excluded from further analyses). ANOVA with Bonferroni corrected posttest: *p < 0.05, **p < 0.01, ***p < 0.001.
… 
Content may be subject to copyright.
1
ScIEnTIFIc RePoRtS | (2018) 8:3825 | DOI:10.1038/s41598-018-22051-z
www.nature.com/scientificreports
New thin-lm surface electrode
array enables brain mapping with
high spatial acuity in rodents
W. S. Konerding1, U. P. Froriep2, A. Kral
1 & P. Baumho1
In neuroscience, single-shank penetrating multi-electrode arrays are standard for sequentially sampling
several cortical sites with high spatial and temporal resolution, with the disadvantage of neuronal
damage. Non-penetrating surface grids used in electrocorticography (ECoG) permit simultaneous
recording of multiple cortical sites, with limited spatial resolution, due to distance to neuronal tissue,
large contact size and high impedances. Here we compared new thin-lm parylene C ECoG grids,
covering the guinea pig primary auditory cortex, with simultaneous recordings from penetrating
electrode array (PEAs), inserted through openings in the grid material. ECoG grid local eld potentials
(LFP) showed higher response thresholds and amplitudes compared to PEAs. They enabled, however,
fast and reliable tonotopic mapping of the auditory cortex (place-frequency slope: 0.7 mm/octave), with
tuning widths similar to PEAs. The ECoG signal correlated best with supragranular layers, exponentially
decreasing with cortical depth. The grids also enabled recording of multi-unit activity (MUA), yielding
several advantages over LFP recordings, including sharper frequency tunings. ECoG rst spike latency
showed highest similarity to supercial PEA contacts and MUA traces maximally correlated with PEA
recordings from the granular layer. These results conrm high quality of the ECoG grid recordings and
the possibility to collect LFP and MUA simultaneously.
An important goal in neuroscience is to understand the activation patterns of neuronal networks at a high spatial
and temporal resolution. is knowledge is important in a variety of medical applications, such as hearing res-
toration via cochlear implants. Both following auditory deprivation and restoration of hearing, spatio-temporal
activation patterns of the auditory pathway are substantially altered1. e characterization of the spatio-temporal
activation patterns in the auditory cortex has commonly been assessed via recordings with penetrating
multi-electrode arrays (MEAs), which have the disadvantage of damaging the brain tissue2. is is especially
disadvantageous for chronic recordings in behaving animals2,3. Dierent MEAs can be used that cover a wide
range of temporal and spatial resolutions (~10 µm to 10 cm; for review see Lebedev & Nicolelis, 2017)4. One ver-
sion of MEAs is the laminar single-shank microelectrode array used for comparison in this study, which allows
conclusions on single cell activity5. In order to track ongoing changes in a chronic setting, for example aer dep-
rivation and following sensory restoration, spatially and temporally precise but non-invasive recording methods
are favorable. Both local eld potentials (LFPs) and action potential related activity are important to understand
the complex neuronal responses to sensory stimulation6. LFPs are commonly assumed to reect the input to the
dendritic eld, generated by synchronized synaptic activity, however, several electrical discharges may add to
these slow potentials (<100 Hz) including Na+ and Ca2+ spikes, ionic uxes through voltage- and ligand-gated
channels, and intrinsic membrane oscillations6,7. e outputs of the respective neuronal tissue are action poten-
tials, recorded extracellularly either as single spikes or as multi-unit activity (MUA), if they are recorded in some
distance from the active cells. If the number of underlying cells cannot be resolved due to small amplitude and/ or
high overlap, this spiking activity (>300 Hz) is sometimes referred to as “hash6. LFP and spiking activity are not
mutually exclusive measures: First, LFPs indicate events that are causal to action potentials8 and secondly, MUA
activity is highly correlated with high-gamma (80–200 Hz) power of the brain oscillations9. It has been shown
that spatial characteristics of cortical responses, such as tonotopy, are represented at a ner spatial resolution by
spiking activity than by slow wave local eld potentials (LFPs) recorded with the same electrodes. is is due to
1Institute of AudioNeuroTechnology and Department of Experimental Otology, ENT Clinics, Stadtfelddamm 34,
Hannover Medical School, 30625, Hannover, Germany. 2Translational Biomedical Engineering, Fraunhofer Institute for
Toxicology and Experimental Medicine (ITEM), Nikolai-Fuchs-Strasse 1, 30625, Hannover, Germany. Correspondence
and requests for materials should be addressed to W.S.K. (email: konerding.wiebke@mh-hannover.de)
Received: 13 November 2017
Accepted: 16 February 2018
Published: xx xx xxxx
OPEN
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
2
ScIEnTIFIc RePoRtS | (2018) 8:3825 | DOI:10.1038/s41598-018-22051-z
the fact that the eect of volume conduction is higher for LFPs and the small-amplitude, high frequency spiking
activity diminishes faster with distance8,1012 (but see Herreras, 2016)13. However, spiking activity is commonly
only assessable using penetrating electrode arrays (PEAs). Subdural surface grids used in electrocorticography
(ECoG) permit simultaneous recording of multiple sites over a large area of the cortex without the need for
repeated (i.e. time consuming) penetrations of the brain tissue. But the combined eects of large distance to the
neuronal tissue, large electrode contact size and high impedances (>1 M at 1 kHz) usually limit surface record-
ings to local eld potentials (LFP) and reduce spatial selectivity (usually >1 mm)4,14.
MUA recorded from the surface of the neocortex is mainly associated with activity of layer 1 interneurons as
well as pyramidal cells and interneurons in deeper layers15. In the study of Khodagoly and colleagues15, isolation
of single neuron action potentials was possible due to the neuron size, electrode diameter (10 µm) and the use of
an organic interface material that conducts not only electric, but also ionic current. However, gold or platinum
contacts are used in conventional ECoG recordings for both humans and animals14,16,17. Downscaled µECoG
electrode grids for human use, with 1-2 mm electrode diameter and -spacing, enable LFP recording which are of a
comparable spatial resolution to those of PEAs (i.e. several hundred µm)14. For animal studies, surface grids with
lower impedances (~200 kOhm) and smaller electrode diameters (150 µm) have been developed18,19. Additionally,
the introduction of thin-lm technology enables intimate electrode-tissue coupling for precise and reliable map-
ping of spatial activation and thus is also well suited for chronic recordings15,18,19. ese have already been shown
to be suitable to detect MUA, in recordings from the basal root ganglia20. Based on positive ndings in cats, but
not humans, the authors conclude that the intimate electrode-tissue coupling is a key factor for surface MUA
recordings and that low impedances and a exible substrate material are critical for high recording quality.
In the present study a thin-film surface ECoG grid was developed in cooperation with Blackrock
Microsystems, Europe, based on the current knowledge in the eld. e grids were then evaluated for their
potential to record spontaneous and evoked cortical activity. As there is currently no standard ECoG electrode
established for animal research, the data was compared to a conventional single-shank PEA (‘Michigan probe’)17.
We found that the new ECoG grids were suitable for recording both LFPs at high spatial resolution and MUA
comparable to simultaneous recordings from PEAs.
Materials and Methods
Animals and surgical preparation. e experiments were performed in 14 (2 female) Dunkin-Hartley
(albino) guinea pigs (372 ± 48 g). All procedures were in accordance with the German and European Union
guidelines for animal welfare (ETS 123, Directive 2010/63/EU) and were approved by the German state authority
(Lower Saxony state oce for consumer protection and food safety, LAVES; approval No. 14/1514). Normal hear-
ing was conrmed by auditory brainstem responses (ABRs; see acoustic stimulation for details).
Anesthesia was induced by an intra muscular injection of a combination of ketamine (50 mg/kg BW), xylazine
(10 mg/kg) and atropine sulfate (0.1 mg/kg). For subsequent inhalation anesthesia, a custom-made endotracheal
tube was inserted through a tracheotomy and connected to a ventilator (Rodent Ventilator 7025, Ugo Basile,
Comerio, Italy). Aer surgical preparation, an adequate anesthesia level for cortical recordings was maintained
by <1.5% isourane in O2/air and was surveyed by testing for paw-withdrawal and corneal reexes. Vital func-
tions were assessed by electro-cardiography (ECG) and capnometry (end-tidal CO2 vol%; Normocap CO2 & O2
Monitor, Datex, Helsinki, Finland). Body core temperature was kept at ~38.0 °C using a heating pad, controlled
via feedback from a rectal probe (TC-1000 Temperature Controller, CWE Inc., Ardmore, USA).
For xation in a stereotaxic frame (Stereotaxic Frame 1430, David Kopf Instruments, Tujunga, USA), the
skull was exposed and a head-holder (custom, stainless steel xation-rod) was secured to the bone anterior to
suture-point Bregma using 3 bone screws (Ø 0.85 mm, Fine Science Tools GmbH, Heidelberg, Germany) covered
with dental acrylic cement (Paladur, Heraeus Kulzer GmbH, Dormagen, Germany). For recordings from the
auditory cortex, a unilateral craniotomy (~5 × 5 mm) was performed, centered at ~2.5 mm caudally from Bregma
and 7.3 mm laterally from the midline. Aer removing the dura mater and positioning of the recording electrodes
the brain was covered with medical grade silicone oil (M 5000, Carl Roth GmbH & Co. KG, Karlsruhe, Germany)
to prevent dehydration through evaporation.
Recording electrodes. e size of the ECoG grid (4 × 4 contacts, 4 mm2 surface area; 0.5 mm contact
spacing) was chosen to cover the average area of the primary auditory cortex (A1) of the guinea pig. e sub-
strate for the grid was a 20 µm thin parylene C lm. e parylene C lm was metallized with gold, which was
photoresist-coated and developed, following direct writing (DWL 66 + , Heidelberg Instruments). Aer etching
of gold and removal of photoresist residuals the grid was connected to a standard Omnetics connector to enable
connection to the recording system. e nal grids had 16 AU contacts (100 µm in diameter) with impedances of
approximately 200 k (see results). Dened openings in the parylene C substrate enabled simultaneous insertion
of a PEA (A1x16-5mm-150-177-A16, NeuroNexus, Ann Arbor, USA). For the characterization of responses to
broad-band stimuli, we recorded simultaneously from 14 cortical positions (12 animals) with 244 recording sites
for the ECoG grid and the PEA, respectively. To assess spatial selectivity in terms of tonotopy, we recorded from
21 surface positions overall (N = 14 animals, N = 336 ECoG recording sites) in combination with 18 positions
from a PEA with 16 channels and impedances of 1.0 ± 0.1 M at 1 kHz (N = 288 recording sites). A silver ball
electrode covered in salt-free electrode gel (Spectra 360, Parker Laboratories INC., New Jersey, USA) was placed
through a hole ~1 mm rostral from Bregma onto the dura mater as a recording reference for both the ECoG grid
and the PEA It was sealed against leakage of cerebro spinal uid with surgical tissue adhesive (Histoacryl, B.
Braun Melsungen AG, Melsungen, Germany) and covered with dental acrylic cement.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
3
ScIEnTIFIc RePoRtS | (2018) 8:3825 | DOI:10.1038/s41598-018-22051-z
Acoustic stimulation. Prior to surgery, ABR click thresholds (i.e. lowest sound intensity inducing a visual
discernable response) were assessed via two trans-dermal silver wire recording electrodes: one at the vertex and
one as retro-auricular reference electrode. e 50 µs condensation clicks were presented via an audiometric head-
phone speaker (DT48, Beyerdynamic, Heilbronn, Germany) in een 5-dB steps (sound pressure level indicated
as peak equivalent: 26–91 dBSPL pe), with 50 repetitions each. Normal hearing was assumed for ABR thresholds of
35 dBSPL pe.
All other acoustic stimuli were presented in randomized order with 30 repetitions each (recording interval:
827 ms). Stimuli were delivered via the audiometric headphone speaker to the ear contralateral to the cortical
recording side. e stimulation was performed in closed eld condition, using a custom build polyvinyl chlo-
ride cone, xed to the speaker and set onto the external meatus. e responsiveness to broad-band stimuli was
accessed using 100 ms white noise bursts (10 ms cosine ramps), presented in seventeen 5-dB steps (sound pressure
level indicated as root mean square value: 0–80 dBSPL rms). To describe the tonotopic organization of A1, pure
tones (100 ms, 10 ms cosine-ramps) at frequencies between 1 kHz and 32 kHz with usually 0.5 octave increments
were presented in nine 10-dB steps (0–80 dBSPL). e sound levels were calibrated prior to the experiments at
the tip of the stimulation cone through the stimulation soware (AudiologyLab, Otoconsult, Frankfurt a.M.,
Gemany) using a condenser microphone (1/4 microphone [4939] in combination with a preamplier [2670] and
a Nexus conditioning amplier [2690], Brüel & Kjaer, Nærum, Denmark).
Data recording and analysis. Recordings of both the ECoG grid and the PEA were performed using a
custom build recording setup and soware (AudiologyLab, Otoconsult, Frankfurt, Germany). e signals were
acquired and amplied through a multichannel recording system (Lynx-8 amplier system, amplication 8000
or 5000 times, butterworth lter: 1 Hz–9 kHz, rollo: 12 dB per octave, Neuralynx, Bozeman, USA) and stored
through AudiologyLab at a sampling rate of 25 kHz using a 32-channel MIO card (NI-6259 National Instruments,
Austin, USA). e three deepest PEA contacts were excluded from the analysis, as they were commonly inserted
beyond the grey matter of the AC due to the length of the PEA. These contacts usually did either show no
responses or did not record activity above the threshold criteria described below.
e peak to peak amplitude of the LFP (resampled at 2 kHz via a Matlab routine) was analyzed from 0–200 ms
aer stimulus onset. MUA spike rates (spikes/stimulus/ms) were measured from 10–40 ms aer stimulus onset.
MUA was calculated as described previously21. In short, the signal was ltered (zero-lack, 2nd order elliptic lter:
300–3000 Hz) and spikes with at least 0.08 ms duration above the detection threshold (i.e. 3* standard deviation,
SD, based on median activity of the signal)22 were counted as MUA. e level of background activity was meas-
ured during a 50 ms pre-stimulus time window and was subtracted from the respective measure (LFP amplitude
or MUA rate) during the post-stimulus time. For comparison of onset-response timing between ECoG grid and
PEA, the rst-spike latency (FSL) was calculated from stimulus onset, as the time when the given spike train dif-
fered signicantly (p < 0.001) from a Poisson distribution23.
As measure of overall responsiveness, we described input-output functions for broad-band noise stimuli.
We excluded data sets as ‘non-responsive’ for which the highest sound intensity did not evoke at least twice the
average background activity, calculated as mean value over all intensities for each contact. We tted the functions
with a sigmoidal t (e.g. Konerding et al., 2017)24 and derived the response threshold (i.e. sound intensity [dB]
inducing 10% of maximal response amplitude) and dynamic range (90% of max–10% of max). In individual
cases, the highest sound intensity (80 dBSPL rms) induced lower response amplitude than the second highest (75
dBSPL rms), potentially due to the middle ear muscle reexes25. In these cases, the tting was performed without
including the responses to 80 dBSPL rms. To gain a comparable measure of response strength for MUA and LFP,
we calculated the response-background ratio (RBR) at 20 dB above response threshold as dB above the average
background activity. As the MUA is mainly a binary measure, we assessed the RBR in terms of changes in spike
rate above background MUA rate26.
e tonotopic organization of the auditory cortex was determined based on the characteristic frequency (CF,
frequency that elicits responses at lowest sound intensities) for each recording site. To calculate the CF, we inter-
polated a tuning curve at 10% of the maximal LFP amplitude or MUA rate of a given contact27, if the maximum
exceeded 3 times the standard deviation of the background activity. e tuning width was calculated in terms of
Q20-values: CF divided by the bandwidth of the tuning curve 20 dB above the sound intensity at CF (from here:
CF threshold). For visual inspection of the tonotopy, we dened CFs with a resolution of 1 octave and plotted all
contacts, spatially aligned based on arterial and suture landmarks. Based on the border of A1 to the dorsal cortex
(DC), as inferred from a frequency reversal, we dened the tonotopic axis orthogonal to the border. To compen-
sate for inter-individual dierences in cortex structure, we normalized the data according to the coordinates of
the 8 kHz CF (taking the mean if there were several sites with 8 kHz for one individual). Subsequently, a linear
regression was calculated along this tonotopic axis to assess the spatial tonotopy resolution in mm/octave, for
each individual and for the whole data set.
e subdural ECoG grid was compared to the PEA with regard to both horizontal and laminar distance.
For the CF distribution along the surface, the PEA contact with lowest CF threshold is compared to the ECoG
grid contacts pooled according to distance from the insertion point: close = 0.5 mm (i.e. 4 inner contacts),
medium = 1 mm, far = 1.3 mm (Fig.1A). For every distance, the CF dierence between ECoG grid and PEA
contacts is calculated in octaves. If several PEA contacts showed the same low CF threshold, the median CF was
chosen (two data sets excluded, as no CF was discernable at the PEA). To estimate the laminar contribution to
the ECoG recordings, we assessed the similarity in FSL and the correlation strength. e FSL was calculated
for the best frequency (BF). e BF was dened as the frequency that elicited highest summed spike rates over
all sound levels. It was used as non-interpolated estimate of the CF (based on MUA), since both measures were
highly correlated (spearman correlation: p < 0.001, r = 0.915 N = 317; Supplement Fig.1). We used the mean FSL
of the four ECoG contacts surrounding the insertion point and pooled the PEA contacts based on the estimated
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
4
ScIEnTIFIc RePoRtS | (2018) 8:3825 | DOI:10.1038/s41598-018-22051-z
location in dierent cortical layers: supragranular layers: 0–600 µm (contact #1–5), granular layer: 800–1100 µm
(contact #6–9), infragranular layers: 1350–1800 µm (contact #10–13). e correlation between the ECoG grid
recordings and the PEA was assessed in response to noise bursts to exclude the inuence of frequency selectivity.
e correlation strength was assessed as two-dimensional (time and repetition) Pearson correlation (r2) between
the 4 ECoG contacts closest to the insertion point and all 16 PEA contacts, using single sweeps (autocorrelation:
r2 = 1). Recordings were excluded whenever clipping was observed in more than 2 individual sweeps. To assess
the correlation of LFPs, we analyzed the raw signals in a time window of 0–200 ms from stimulus onset. To corre-
late spiking activity at the surface and in the depth we transferred the MUA event train into a continuous signal by
calculating the moving average (1 ms integration width). As an estimate of overall correlation between the ECoG
grid and the PEA, we calculated a sweep-wise sum over all 16 PEA contacts. e surface contact with the highest
correlation to this summed signal was used to analyze the eect of insertion depth, correlating each of the 16
channels, separately, with the respective ECoG contact.
Figure 1. Recordings from the newly developed ECoG grid were possible for all 16 surface contacts and
were similar to simultaneous recordings from a conventional penetrating multi-electrode array. (A) Sketch of
the subdural ECoG grid. Markers on the grid facilitate the allocation to the respective surface locations. e
distances to the central insertion point are indicated (a = 0.5 mm, b = 1.0 mm, c = 1.3 mm) (B) e penetrating
multi-electrode array (PEA) is inserted through an opening in the ECoG grid substrate. Marked are the pseudo
sylvian sulcus (PSS) and the middle cerebral artery (MCA), which served as landmarks for spatial alignment
of dierent recording positions. (C) Example image of a 16 channel single shank PEA. (D) Examples of traces
(raw signal) for an ECoG grid and a penetrating multi-electrode array from the same cortical recording during
spontaneous activity.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
5
ScIEnTIFIc RePoRtS | (2018) 8:3825 | DOI:10.1038/s41598-018-22051-z
Statistical analysis. Considering the focus of this study and due to the difference in electrode con-
tact characteristics (e.g. size and impedance), we assessed the recordings of the ECoG grid and PEA contacts
as independent measures and performed unpaired analyzes for the comparison of LFP characteristics at both
electrodes. We performed paired comparisons for LFP and MUA recorded at the same electrode contact. e
Kolmogorov-Smirnov test was used to conrm normal distribution of the data and, if required, non-parametric
analyzes were applied and the median ± interquartile range (IQR) instead of the mean ± standard deviation (SD)
was calculated. e level of signicance was set to 5%.
Results
We rst described the properties of the new ECoG grid and its potential for recording spontaneous activity
from the auditory cortex. Subsequently, we showed the similarity of LFP recordings from the ECoG grid to
simultaneously recorded LFP of PEAs. We assessed both the evoked responses to broad-band stimuli in terms of
input-output functions and showed stimulus specicity in terms of tonotopy, using tonal stimulation. To explore
the origin of the ECoG recordings, we furthermore correlated subdural recordings with recordings from dierent
cortical depths. Finally, we described the potential of the ECoG grid to record MUA from the surface and related
the derived measures to those of the LFP recordings.
ECoG grid design and spontaneous activity recording. e grid substrate parylene C achieved high
exibility and enabled an intimate contact at the electrode-tissue interface. e average impedance was 212 ± 51
k (measured at 1 kHz; n = 5 grids), ranging from 128 to 395 k (one extreme value with 962 k). Individual
ECoG grids were used up to six times, without observable changes in recording quality. Simultaneous recordings
from a PEA inserted through one of the openings in the grid material were possible with high signal quality
(Fig.1).
Evoked LFP responses to broad-band stimuli. To assess the potential of the new ECoG grid to record
brain activity during sensory stimulation (Fig.2A), we analyzed the evoked LFP responses to broad-band stimuli
recorded with the ECoG grid and compared these to simultaneous recordings from PEAs at 14 cortical positions.
Of 224 recording sites, 126 on the ECoG grid showed a clear onset response (56.3%; criterion: >2* background
activity) while this was the case for 58 (31.9%) recording sites on the PEA (N = 182; 3 deepest contacts excluded).
e input-output functions (Fig.2B) showed some variability across recording sites (Fig.2B), and the sigmoi-
dal t was not possible for 15 ECoG and 4 PEA responsive sites. In the remaining data sets, the goodness of
t was high (ECoG: r2 = 0.956, SD: 0.046, n = 111; PEA: r2 = 0.955, SD: 0.025, n = 54). e calculated maximal
peak to peak amplitude was signicantly higher for the ECoG (median: 1.177 mV) as compared to the PEA
(0.478 mV; Mann-Whitney U test: p < 0.0001, U = 1475; Fig.2C). e response-background ratio at 20 dB above
response threshold was higher for the ECoG grid (median: 0.455, n = 101, negative values excluded) than for the
PEA (median: 0.292, n = 52, negative values excluded; Mann-Whitney U test: p < 0.0001, U = 1308; Fig.2D).
e calculated response threshold was higher at ECoG contacts (median: 32.46 dBSPL rms) than at PEA contacts
(21.75 dBSPL rms; Mann-Whitney U-test: p < 0.0001, U = 1817; Fig.2E). e dynamic range was however similar
for ECoG grid and PEA contacts (medianECoG: 29.79 dB; medianPEA: 36.10 dB; Mann-Whitney U test: p = 0.369,
U = 2738).
Tonotopy based on LFP recordings. To determine the spatial selectivity of recordings from the ECoG
grid, we assessed the frequency selectivity of each recording site and the change in characteristic frequency
(CF) with cortical position (i.e. tonotopy). For each recording site at the ECoG grid (N = 336) and at the PEA
(N = 234), we determined the CF and the tuning width in terms of Q20-values from the frequency response
curves based on LFP recordings (Fig.3A). Based on the dened CF threshold (criterion see methods), 12% of
the ECoG recording sites were dened unresponsive and were excluded from the analysis. In another 10% no
single CF could be determined, due to the double-peaked appearance of the frequency tuning curve (FTC).
Additionally, there was 1 case in which the frequency response area showed no clear FTC and in 15 cases the CF
was at or possibly below 1 kHz (lowest stimulated frequency) and no Q20-value could be assessed. Of the PEA
recordings, 41% were dened unresponsive (i.e. fell below CF threshold criterion), 13% of the FTCs were double
peaked, 3% had a CF at 1 kHz and 1 case (not the same animal as for the ECoG grid) had a complex response
characteristic.
e average Q20 tuning width was similar for the surface and penetrating electrodes (medianECoG = 0.995,
medianPEA = 0.916; Mann-Whitney U test: p = 0.361, U = 11411, nECoG = 246, nPEA = 99). is dierence was not
based on a dierences in CF, as ECoG and PEA contacts had similar CF values (medianECOG = 8.1, medianPEA =
8.0; p = 0.797, U = 26381, nECoG = 261, nPEA = 205). e spatial distribution of CF of the combined ECoG record-
ing sites showed a frequency reversal, which marks the transition from the primary auditory cortex (A1) to the
dorsocaudal cortex (DC, Fig.3B).
Based on a virtual line of the medio-lateral frequency reversal we dened the orthogonal, rostro-caudal tono-
topic axis. All CFs within A1 were correlated with their cortical coordinates relative to the tonotopic axis and for
each individual we derived the slope of the resulting linear regression (Fig.3C). One animal was excluded due to
very poor tting (r2 = 0.057), the remaining 13 datasets had a high coecient of determination (r2 = 0.709, SD:
0.139). e slope for all data was 0.678 mm/octave (r2 = 0.685). e individual slopes had a low spread around
the mean (0.646 mm/octave) with a standard deviation of 0.153 mm/octave and a range from 0.373 to 1.000 mm/
octave.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
6
ScIEnTIFIc RePoRtS | (2018) 8:3825 | DOI:10.1038/s41598-018-22051-z
Correspondingly, the CF dierence between the ECoG and PEA contacts was around 1 octave for contacts
close (0.5 mm) to the insertion point (mean ± SD: 0.918 ± 0.869 octave) and signicantly increased with horizon-
tal distance (repeated measure 1-way ANOVA: p < 0.0001, n = 16; Fig.3D).
Correlation with laminar distance. In order to determine the origin of the LFP signals recorded at the ECoG
grid, we correlated the signal with those recorded via the PEA. e correlation strength (r2) signicantly declined
with cortical depth from channel #1 to #16, with a high correlation for the most supercial channel #1 (r2 = 0.411)
and weakest correlation for channel #16 (r2 = 0.036; 1-way repeated measure ANOVA: p < 0.0001, F = 8.426,
df = 15; Fig.4). e comparison with the correlation strength between ECoG grid and summed signal of the PEA
(r2 = 0.190) revealed that only the signal recorded at the rst contact showed a signicant higher correlation with
the ECoG grid recordings. e summed signal of the PEA explained less variance of the ECoG recordings than
each of the 4 supercial PEA contacts, individually.
Figure 2. e new ECoG grid enabled recordings of evoked LFP responses similar to conventional penetrating
multi-electrode arrays. (A) Examples of evoked responses to a 100 ms broad-band stimulus for an ECoG grid
and a penetrating multi-electrode array (PEA) from the same cortical recording. Given are 30 single sweeps
(thin grey lines) and an averaged signal (thick black line). Stimulus onset is indicated (red arrow head). (B)
Input-output functions of LFP peak-to-peak (p2p) amplitudes in response to a broad-band stimulus. e colors
represent dierent recording positions; the three deepest PEA contacts (excluded from further analysis) are
indicated in light grey. e amplitudes at the penetrating electrode (PEA) are signicantly smaller than those
at the ECoG grid (C and D). (E) e response thresholds at ECoG grid contacts was slightly higher (i.e. worse)
than those recorded at the penetrating electrode (PEA) contacts. (C) e maximal peak to peak (p2p) amplitude
at ECoG grid contacts was signicantly higher than those recorded at the penetrating electrode (PEA) contacts.
(D) e response-background ratio (RBR) at 20 dB above response threshold was signicantly higher at ECoG
than at PEA contacts. (CE) Given are individual data (dots) and medians with IQR (lines). Mann-Whitney U
test: ***p < 0.0001.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
7
ScIEnTIFIc RePoRtS | (2018) 8:3825 | DOI:10.1038/s41598-018-22051-z
Comparison of MUA activity at the ECoG grid to the PEA. We analyzed the high frequency component (300–
3000 Hz) of the recorded signal that appeared as hash in the single traces (Fig.5). From 224 and 182 record-
ing sites at the ECoG grid and the PEA, respectively, 193 (86.2%) and 83 (45.6%) were dened responsive to
broad-band stimuli based on a MUA rate (Fig.5) of at least twice the background activity. e MUA had a clear
onset response with a sigmoidal input-output function (Fig.5B). Usually, also a weak oset response was discern-
able (Fig.5C, data not analyzed).
To determine the potential origin of MUA recorded at the ECoG grid, we compared the FSL between ECoG
and PEA contacts and also derived the correlation strength between the respective signals. The mean FSL
(22.65 ms) of the four surface contacts surrounding the insertion point was similar to those recorded by super-
cial contacts of the PEA (depth 1/ supragranular layer: 24.33 ms, depth 2/ granular layer: 26.27 ms) and diered
signicantly from the one in deeper, infragranular layers of the cortex (depth 3: 27.45 ms; dependent t-tests with
Bonferroni correction: surface vs depth1 p = 0.792; surface vs depth2 p = 0.342; surface vs depth3 p = 0.012,
N = 15, n = 3 comparisons; Fig.5C).
e correlation strength of the MUA between the ECoG and the PEA contacts showed signicant changes
with cortical depth (ANOVA: p < 0.001, F = 4.422, df = 15; Fig.5D). e summed signal over all PEA contacts
yielded a signicantly higher correlation strength with the ECoG grid (r2 = 0.191) than most of the individual
Figure 3. e ECoG grid enabled ne-scale tonotopic mapping of the GP auditory cortex based on LFP
recordings. (A) Example of frequency response curves derived by ECoG grid recordings and recordings from
penetrating electrodes. Indicated are the CF and the lower and upper value 20 dB above CF, from which Q20-
values were calculated. (B) Tonotopic map of the auditory cortex. For each ECoG grid contact (NLFP = 241) the
CF in octave increments was plotted over the respective cortical coordinate. A frequency reversal (dashed line)
marks the transition from the primary auditory (A1) to the dorsocaudal cortex (DC). (C) Correlation of CF
with the tonotopic axis of A1 for LFP recordings from all ECoG grid recording sites. Given are individual data
(dots) and regression lines per individual (gray, thin lines), as well as the overall regression (thick solid line)
with the 95% condence interval (dashed lines). (D) e CF distance in octaves between ECoG grid recording
sites and penetrating multi-electrode array (PEA) increased with increasing surface distance. Given are
individual data (dots) and mean with SD (lines). Repeated measure 1-way ANOVA p < 0.002 with Bonferroni
post-test: ***p < 0.001.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
8
ScIEnTIFIc RePoRtS | (2018) 8:3825 | DOI:10.1038/s41598-018-22051-z
contacts. Only the correlation strength between the ECoG grid and the 7th contact (~900 µm deep, r2 = 0.142) was
not signicantly dierent from the one with the summed signal and also had signicant higher values than most
of the electrode contacts inserted deeper into the cortex. us, the summed signal of the PEA did not explain
signicantly more variance in the surface recordings than the penetrating contact #7 on its own.
Comparison between MUA and LFP measures at the same ECoG grid and PEA contacts. To assess whether the
MUA assessment gave information in addition to the LFP measures, we rst compared the activity to broad-band
stimuli with regard to the response threshold, dynamic range and the response-background ratio (based on LFP
amplitude or MUA rate) at 20 dB above response threshold. e sigmoidal t of the MUA input-output func-
tions was not possible for all contacts. e respective sample sizes for paired data (i.e. LFP and MUA at the
same contact) are given below. e results were not specic to the ECoG grid, but were also found for the PEA.
e response-background ratio was signicantly higher for MUA (ECoG: 1.589, PEA: 1.122) compared to LFP
measures (ECoG: 0.451, PEA: 0.209; Wilcoxon test: ECoG: p < 0.0001, W = 4465, n = 94; PEA: p < 0.0001,
W = 276, n = 23; Fig.6A). e response threshold was signicantly higher for MUA (ECoG: 35.87 dBSPL rms,
PEA: 30.49 dBSPL rms) compared to LFP measures (ECoG: 32.40 dBSPL rms, PEA: 23.29 dBSPL rms; Wilcoxon test:
ECoG: p < 0.0001, W = 37490, n = 94; PEA: p = 0.0002, W = 242, n = 23; Fig .6B). e dynamic range was
signicantly lower for MUA (ECoG: 21.58 dB, PEA: 26.62 dB) compared to the LFP (ECoG: 30.27 dB, PEA:
39.18 dB; Wilcoxon test: ECoG: p < 0.0001, W = 3049, n = 102; PEA: p = 0.029, W = 154, n = 24; Fig.6C).
To assess whether the spatial selectivity of the ECoG grid MUA recordings was higher than the one derived
from LFP recordings, we compared the frequency selectivity measures CF slope along the tonotopic axis and the
Q20-values. e slope for all MUA data was 0.712 mm/octave and individual slopes (min: 0.419, max: 1.890 mm/
octave) did not dier signicantly from those derived by LFP measures (Wilcoxon signed rank test: p = 0.588,
W = 17, n = 13 animals; Fig.6D). As for the LFP recordings, the CF dierence between the ECoG and PEA
was below 1 octave for contacts close (0.5 mm) to the insertion point (mean ± SD: 0.937 ± 0.710 octave) and
signicantly increased with increasing distance (repeated measure 1-way ANOVA: p = 0.010, n = 16). e MUA
Q20 tuning width was similar for ECoG and penetrating electrodes (medianECoG = 1.043, medianPEA = 1.073;
p = 0.343, U = 10255, nECoG = 239, nPEA = 92). However, when comparing the LFP tuning widths to those of the
MUA, the MUA Q20-values (ECoG: 1.053, PEA: 1.075) were signicantly larger than the LFP Q20-values (ECoG:
0.979, PEA: 0.823), indicating a slightly higher spatial resolution derived by MUA. is was the case both for
the PEA (Wilcoxon signed rank test: p = 0.009, W = 663, n = 55) and for the ECoG (p = 0.040, W = 2918,
n = 206; Fig.6E). is dierence in tuning width was not based on a dierence in CF (with higher CF having
higher Q-values), as the corresponding CF-values were similar or even lower for MUA than for LFP measures
(ECoG: medianLFP: 8.1, medianMUA : 8.1, p = 0.816, W = 229, n = 220; PEA: medianLFP: 6.8, medianMUA: 2.4,
p < 0.001, W = 1366, n = 71).
Discussion
Subdural ECoG grid recordings allow scanning large brain areas by using multiple recording sites, without dam-
aging the underlying brain tissue. e newly developed ECoG grid (Blackrock Microsystems Europe) was found
to yield similar recording quality as a conventional PEA (i.e. Neuronexus ‘Michigan probe’). e impedances of
the ECoG grids were low (mean 212 k) and showed little variability between contacts (SD: 51 k; one extreme
at 962 k). e thin-lm material allowed intimate contact to the brain surface and recordings of spontaneous
brain activity could be obtained from all ECoG grid contacts. Recording quality was not compromised by cover-
ing the grid and brain by silicone oil. Evoked responses could be reliably assessed in a vast majority of the record-
ing sites and the incidence of non-responsive sites was much lower than for the PEA for both broad-band stimuli
Figure 4. Correlation strength (r2) between grid and penetrating electrode recordings for LFP. Given are mean
(dot) and SEM (whisker) and the non-linear regression line (One-phase exponential decay: r2 = 0.370). Base d
on the raw signal, we revealed that recordings of the grid and the penetrating (PEA) contacts signicantly
declined with cortical depth (contacts #14–16, indicated in light grey, were excluded from further analyses).
Beyond 5 contacts (~ 750 µm), r2 fell below the summed sweep-by-sweep correlation strength over all 16
penetrating contacts (red). ANOVA with Bonferroni corrected posttest: **p < 0.01.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
9
ScIEnTIFIc RePoRtS | (2018) 8:3825 | DOI:10.1038/s41598-018-22051-z
(26 vs. 54%) and pure tones (13 vs. 47%). Although we did not systematically assess the stability of the ECoG grid
over several recording sessions, we observed persisting high recording quality for up to six experiments, using
the same ECoG grid. We found that these new ECoG grids were suitable for recording both LFPs at high spatial
resolution and MUA comparable to conventional PEAs.
By recording simultaneously from the surface ECoG grid and the PEA, we could directly compare evoked
LFP responses between the two electrode types. The ECoG grid LFPs were on average larger, had a higher
response-background ratio and showed slightly elevated response thresholds as compared to the PEA recordings.
e dynamic range was similar between the electrode types.
Figure 5. e newly developed ECoG grid enabled recordings of evoked MUA similar to conventional
penetrating multi-electrode arrays. (A) Examples of ltered (300–3000 Hz) traces with detected spikes (red
circles) above detection threshold (dashed line) for both ECoG and penetrating multi-electrode array (PEA).
Stimulus onset is indicated (red arrow head). (B) Representative examples of averaged LFP and raster-plots
of MUA in response to a broad-band noise stimulus of 80 dBSPL rms. e raster-plots indicate the time points
of every detected spike in each of the 30 repetitions. Stimulus onset is indicated (vertical line). (C) e peri-
stimulus time histograms (psth) show representative responses to a 100 ms broad-band stimulus, for the ECoG
grid and the penetrating electrode (PEA), respectively. Usually a strong onset response and a weak oset
response were discernable. (D) Similarity between rst spike latency (FSL) for ECoG grid and penetrating
multi-electrode array contacts decreases with cortical depth (depth3 = infragranular layer). Given are median
FSLs at the best frequency threshold level (i.e. 10% of max MUA, comparable to the CF threshold) for ECoG
grid recording sites close to the insertion point and for dierent depths at the PEA (depth 1: 0–600 µm, depth
2: 750–1200 µm, depth 3: 1350–1800 µm). Box-plots with min and max values (whisker) and mean (cross). e
FSL at the surface was signicantly shorter than the one recorded in deep layers of the AC; dependent t-tests
with Bonferroni correction *p < 0.05. (E) Correlation strength (r2) between grid and penetrating electrode
recordings for MUA is highest for granular layers (contact #7). Given are mean (dot) and SEM (whisker). Aer
an initial decline in correlation strength, the maximal correlation was reached at the 7th contact and sharply
declined with increasing cortical depth (contacts #14–16, indicated in light grey, were excluded from further
analyses). ANOVA with Bonferroni corrected posttest: *p < 0.05, **p < 0.01, ***p < 0.001.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
10
ScIEnTIFIc RePoRtS | (2018) 8:3825 | DOI:10.1038/s41598-018-22051-z
One potential explanation for signicant dierences between surface and penetrating electrode recordings
would be dierences in cortical depth. However, we did not nd any indications in our data (analyzes not shown)
and to the best of our knowledge, no such dierence have been reported in the literature28. e dierence in
response threshold was unexpected and the distribution of values highly overlapped between the two electrode
types. A dierent involvement of input sources, including long latency activity, may have led to a higher variance
in individual surface recordings and the observed small increase of ECoG grid response thresholds in the aver-
aged signal, as compared to the depth recordings. Larger amplitudes at the ECoG grid contacts as compared to the
PEA may arise due to lower impedances (200 k vs 1 M). However, the inuence of impedance values and elec-
trode contact geometry is discussed to be rather negligible for LFP recordings29. Additionally, larger amplitudes
and higher response-background ratios may arise from recordings from a higher number of synchronously active
neurons7. is would indicate that integration occurs over more neurons at the ECoG grid electrode as compared
to the penetrating electrode contacts. However, the high spatial acuity and similarity in tuning width (see below)
Figure 6. e characterization of the surface responses diered signicantly between MUA and LFP
measures at the same contacts. (A) e response-background ratio (RBR) 20 dB above response threshold was
signicantly higher (i.e. better) for MUA (based on rate) as compared to LFP (based on amplitude). Given are
individual data (dots) and medians with IQR (lines). Wilcoxon test: ***p < 0.0001. (B) e response threshold
was signicantly higher (i.e. worse) for MUA as compared to LFP measures. Given are individual data (dots)
and medians with IQR (lines). Wilcoxon test: ***p < 0.0001. (C) e dynamic range was signicantly lower
for MUA as compared to LFP measures. Given are individual data (dots) and medians with IQR (lines).
Wilcoxon test: *p < 0.05, ***p < 0.0001. (D) e MUA Q20-values were signicantly larger (i.e. sharper tuning)
compared to LFP Q20-values. Given are individual data (dots) and median with IQR (lines). Wilcoxon test:
*p < 0.05, ***p < 0.0001. (E) e tonotopy derived by MUA measures (black) at the ECoG grid was similar to
the one derived by LFP measures (light orange). Given are individual data (dots) and the linear regression with
the 95% condence interval (solid and dashed lines).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
11
ScIEnTIFIc RePoRtS | (2018) 8:3825 | DOI:10.1038/s41598-018-22051-z
showed that the ECoG grid contacts did not integrate over a signicantly larger amount of cells, at least in the
horizontal plain. Furthermore, the size of LFP amplitudes is discussed to be inuenced by the cytoarchitectonic
structure rather than by functional factors13. e fact that the surface electrode is recording from the outside of
the activated tissue, in a dierent orientation to the LFP source(s), as compared to the PEA, may therefore explain
the signicant dierences in LFP magnitude.
e ECoG grid enabled fast and reliable assessment of the tonotopic map, revealing the typical tonotopic
organization of the auditory cortex with a frequency reversal dening the transition between A1 and DC. Most
of the derived frequency tuning curves (FTC) were V-shaped and allowed the designation of a CF and calcu-
lation of Q20-values. Only in about 10% of both the surface and the depth recordings double-peaked FTC30
were found. is corresponds well to observations of uniform V-shapes under anesthesia with increasing num-
bers of multi-peaked FTC in awake animals, e.g. 20% of neurons in marmosets (for review see Schreiner et al.,
2011)31. e individual frequency slopes along the tonotopic axis were highly consistent and the average slope was
0.68 mm/octave (0.71 mm/octave for MUA). is value is similar to the one reported by Hellweg and colleagues
of 0.63 mm/oct ave32 and corresponds well to frequency distributions in A1 reported in the literature15,30,33. ese
were derived by multiple insertions of penetrating electrodes30,32,33 or by optical imaging methods15. Based on
the guinea pigs broad hearing range of 0.05–50 kHz (~9.5 octaves)34 and the size of A1 of 4 mm along the tono-
topic axis (Fig.3C and30), a slope of 0.42 mm/octave would be required to evenly represent the whole frequency
range. is theoretical slope is considerably steeper than the values derived by mapping the auditory cortex. e
discrepancy may be based on under-representation of parts of the hearing range, or omissions of specic frequen-
cies. e two eects have been described in several mammalian species and are both discussed to reect specic
environmental adaptations31. In our data, the distribution of frequencies shows an obvious gap at around 4 kHz,
which has also been shown by previous recordings from the guinea pig auditory cortex12. We assume that this gap
is based on an adaptive response (either plastic change or inborn characteristic) to the resonance frequency of the
bulla, which is at around 4 kHz35. is frequency would dominate the auditory percept, if it was not dampened
or ltered out. is process is most probably established from the cochlea onwards, as an underrepresentation of
CFs around 4 kHz is already apparent in auditory nerve recordings36. A comparison of the ECoG grid tonotopy
to one derived by PEAs was not possible in the present study, as we did only use one penetration per surface
recording position to prevent severe brain damage due to extensive sampling. Besides of the reported similarity to
the tonotopic organization in guinea pig A1 derived by dierent recording methods (see above), we furthermore
compare the Q20-values of the surface and penetrating electrode and found no signicant dierence in sharpness
of tuning.
A detailed analysis of the source and spread of the LFP recordings was not within the scope of the study. We
did however assess the correlation between ECoG and PEA recordings. e results showed that the signal at the
surface was most similar to the ones in supercial layers and the similarity exponentially decreased with cortical
depths. is, together with the high (spatial) frequency selectivity, indicates that the LFP recordings in our study
are quite local. Similar changes in the LFP signal over only few 100 µm have been reported in the literature29. e
actual spread may however be less specic for the type of recording electrode and be rather a characteristic of the
recorded brain structure and sensory modality13.
MUA at the ECoG contacts was similar to the one recorded at the PEA, with similar psth response pro-
les and input-output characteristics. When comparing MUA to LFP recordings, MUA yielded higher response
thresholds, lower dynamic ranges and higher response-background ratios (based on MUA rate)26. e tonotopic
slope was similar to the one derived from LFP measures. However, the Q20-values were slightly higher (sharper
tunings) for the MUA as compared to the LFP recordings. All these dierences were not specic for the surface
ECoG grid, but were also obvious at the PEA, conrming the general dierence between MUA and LFP record-
ings. Most of the results are in line with the notion that MUA is spatially more selective, integrating activity over a
smaller sample of neurons, than the LFP recordings8. e higher spatial selectivity will additionally be sharpened
by lower volume conduction for MUA as compared to LFP, with less overlap between sources12 and by dierences
in the characteristics of the sources13. As the LFP is dominated by the cortical input, whereas the MUA corre-
sponds to the output of the respective neurons, lower frequency selectivity in sub-cortical as compared to cortical
structures will aid to the observed dierence in frequency selectivity37. e elevated response thresholds for MUA
(see also Norena & Eggermont, 2002)37 can also be interpreted this way, considering that at response threshold the
synchrony of activation is important for detection and that variance in a small set of neurons (MUA) will lead to
higher uncertainty than the same variance in a larger set of neurons (LFP). Additionally, LFPs may be inuenced
by cortical input that is too small in amplitude to generate local MUA37. e higher response-background ratio at
20 dB above response threshold is likely to be based on dierences in background activity for LFP amplitudes as
compared to MUA rates. LFP recordings, integrating over a larger population of neurons, will result in a higher
background activity than MUA that is based on a smaller set of neurons. However, due to the dierence in the
stimulus characteristics (eld vs. spikes), a direct comparison between the levels of background activity was not
possible.
e fact that we revealed very similar slopes for CF-changes along the tonotopic axis for both LFP and MUA
measures seemed at rst sight to be contradictory to the assumption of a higher spatial acuity for MUA record-
ings. For example, Fallon and colleagues17 reported that they needed to calculate the second spatial derivative
(SSD) to reveal similar tonotopic slopes of surface LFP recordings compared to MUA PEA recordings. However,
for the visual cortex (V1) of macaques, Xing and colleagues38 found similar visual eld maps for simultaneously
recorded LFP and MUA at the same penetrating electrode and CF values have typically been reported to be
similar between LFP and MUA PEA recordings12,37. We assume that the contact size and distance of the (surface
or penetrating) recording electrodes to the respective neurons, in relation to their functional selectivity (e.g.
frequency tuning), is determining whether or not the higher spatial acuity of MUA recordings will result in sig-
nicant dierences of local characteristics (e.g. CF), as compared to LFP recordings.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
12
ScIEnTIFIc RePoRtS | (2018) 8:3825 | DOI:10.1038/s41598-018-22051-z
To assess the link between ECoG and PEA MUA we rst compared the FSL and revealed similar FSL between
the ECoG grid recordings and recordings from supragranular layers. Furthermore, the MUA trains at the surface
were best correlated with trains recorded at the granular layer IV (800–1100 µm)28. As the granular layer showed
higher spiking activity than supragranular layers (Fig.5A), this may be an artifact of the correlation computation,
as lower correlation coecients may result from less variability in the data (i.e. few spikes)39. However, adding the
signals of all 16 contacts did not result in further signicant improvement of the correlation. us, we conclude
that the MUA recorded at the ECoG grid corresponds to supragranular and granular (~900 µm depth) spiking
activity. It is however unlikely that an individual granular spike is picked up at a subdural ECoG contact, as the
horizontal spread of MUA is supposed to be around 200 µm6,15,40,41. We assume that the MUA (or hash) recorded
at the surface corresponds to a summation of simultaneous spiking activity40 of layer 2/3 pyramidal cells, which
may also include activity from layer 1 interneurons15 and backwards propagating, dendritic spikes from deeper
layers42. Further research, such as intracortical electrical stimulation43 and pharmacological blockage, will be
necessary to characterize the origin of subdural spiking activity recorded with the ECoG grid.
e fact that the newly developed surface electrode also enabled recordings of MUA was quite unexpected, as
previous publications usually do not document this feature (for review see Im & Seo, 2016)44, or link it to specic
properties of the electrode, such as organic interface materials15. However, gold and platinum ECoGs have already
been described to be able to record MUA from basal root ganglia in the cat, but not in humans, and the authors20
conclude that the distance between electrode and tissue might be the key issue for recording unit-activity and
suggest that reduced impedances and exible substrate material would improve the recording quality. e ECoG
grid in our study addressed both and provided a exible and adhesive substrate with low impedance contacts.
Due to the lack of systematic analysis of the potential to record surface MUA in current publications (probably
related to a reluctance to report negative results), we cannot estimate which material characteristics of the ECoG
are critical for this feature. e MUA recording yields several advantages as compared to sole LFP recordings.
In addition to the higher spatial acuity, the MUA is also known to allow assessing temporal aspects with higher
delity. For example, phase locking to repetitions within the signal above 16 Hz is possible, which is the limit for
LFP frequency following responses45. By this, the MUA is able to reect for example the fundamental frequency
of speech46.
Conclusion
Taken together, the study revealed that the new ECoG grid is a good alternative to conventional PEA considering
the comparable signal quality and the potential to record LFP as well as spatially selective MUA. Additionally,
they provide the advantage of faster sampling from multiple cortical sites simultaneously, without the risk of tis-
sue damage. e recordings can easily be combined with PEA recordings, whenever layer specic information is
needed. e next step would be biocompatibility testing during chronic preparations. Based on the intimate con-
tact to the brain tissue and the stability over several recording sessions, we expect the ECoG grid to be a relevant
tool for chronic recordings in the behaving animal.
Dataset availability. e datasets generated during and/or analyzed during the current study are available
from the corresponding author on reasonable request.
References
1. ral, A. et al. Spatiotemporal patterns of cortical activity with bilateral cochlear implants in congenital deafness. Journal of
Neuroscience 29, 811–827 (2009).
2. Ward, M. P., ajdev, P., El lison, C. & Irazoqui, P. P. Toward a comparison of microelectrodes for acute and chronic recordings. Brain
Research 1282, 183–200, https://doi.org/10.1016/j.brainres.2009.05.052 (2009).
3. Biran, ., Martin, D. C. & Tresco, P. A. Neuronal cell loss accompanies the brain tissue response to chronically implanted silicon
microelectrode arrays. Experimental neurology 195, 115–126 (2005).
4. Lebedev, M. A. & Nicolelis, M. A. L. Brain-Machine Interfaces: From Basic Science to Neuroprostheses and Neurorehabilitation.
Physiological Reviews 97, 767–837, https://doi.org/10.1152/physrev.00027.2016 (2017).
5. Meredith, M. A. & Allman, B. L. Single-unit Analysis of Somatosensory Processing in Core Auditory Cortex of Hearing Ferrets. e
European journal of neuroscience 41, 686–698, https://doi.org/10.1111/ejn.12828 (2015).
6. Bulloc, T. H. Signals and signs in the nervous system: the dynamic anatomy of electrical activity is probably information-rich.
Proceedings of the National Academy of Sciences 94, 1–6 (1997).
7. Buzsái, G., Anastassiou, C. A. & och, C. e origin of extracellular elds and currents-EEG, ECoG, LFP and spies. Nature
reviews neuroscience 13, 407–420 (2012).
8. ajiawa, Y. & Schroeder Charles E. How Local Is the Local Field Potential? Neuron 72, 847–858, https://doi.org/10.1016/j.
neuron.2011.09.029 (2011).
9. ay, S. & Maunsell, J. H. . Dierent origins of gamma rhythm and high-gamma activity in macaque visual cortex. PLoS Biol 9,
e1000610 (2011).
10. Eggermont, J. J., Munguia, ., Pienowsi, M. & Shaw, G. Comparison of LFP-Based and Spie-Based Spectro-Temporal eceptive
Fields and Cross-Correlation in Cat Primary Auditory Cortex. PLOS ONE 6, e20046 (2012).
11. Gaucher, Q., Edeline, J.-M. & Gourévitch, B. How dierent are the local eld potentials and spiing activities? Insights from multi-
electrodes arrays. Journal of Physiology-Paris 106, 93–103 (2012).
12. de Cheveigné, A., Edeline, J.-M., Gaucher, Q. & Gourévitch, B. Component analysis reveals sharp tuning of the local eld potential
in the guinea pig auditory cortex. Journal of Neurophysiology 109, 261–272, https://doi.org/10.1152/jn.00040.2012 (2013).
13. Herreras, O. Local Field Potentials: Myths and Misunderstandings. Frontiers in Neural Circuits 10, 101, https://doi.org/10.3389/
fncir.2016.00101 (2016).
14. ellis, S. et al. Multi-scale analysis of neural activity in humans: Implications for micro-scale electrocorticography. Clinical
Neurophysiology 127, 591–601, https://doi.org/10.1016/j.clinph.2015.06.002 (2016).
15. hodagholy, D. e t al. NeuroGrid: recording action potentials from the surface of the brain. Nature neuroscience 18, 310–315 (2015).
16. Toda, H. et al. Simultaneous recording of ECoG and intracortical neuronal activity using a exible multichannel electrode-mesh in
visual cortex. NeuroImage 54, 203–212, https://doi.org/10.1016/j.neuroimage.2010.08.003 (2011).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
13
ScIEnTIFIc RePoRtS | (2018) 8:3825 | DOI:10.1038/s41598-018-22051-z
17. Fallon, J. B. et al. Second spatial derivative analysis of cortical surface potentials recorded in cat primary auditory cortex using thin
lm surface arrays: Comparisons with multi-unit data. Journal of neuroscience methods 267, 14–20, https://doi.org/10.1016/j.
jneumeth.2016.04.004 (2016).
18. Hollenberg, B. A., ichards, C. D., ichards, ., Bahr, D. F. & ector, D. M. A MEMS fabricated exible electrode array for recording
surface eld potentials. Journal of neuroscience methods 153, 147–153 (2006).
19. Hosp, J. A. et al. Thin-film epidural microelectrode arrays for somatosensory and motor cortex mapping in rat. Journal of
neuroscience methods 172, 255–262, https://doi.org/10.1016/j.jneumeth.2008.05.010 (2008).
20. Gaunt, . A. et al. Single- and Multi-Unit Activity ecorded from the Surface of the Dorsal oot Ganglia with Non-Penetrating
Electrode Arrays. Conference proceedings: Annual International Conference of the IEEE Engineering in Medicine and Biology Society.
IEEE Engineering in Medicine and Biology Society. Conference 2011, 6713–6716, https://doi.org/10.1109/iembs.2011.6091655 (2011).
21. Tillein, J., Huba, P. & ral, A. Monaural congenital deafness aects aural dominance and degrades binaural processing. Cerebral
Cortex 26, 1762–1777 (2016).
22. Quian Quiroga, ., Nadasdy, Z. & Ben-Shaul, Y. Unsupervised spie detection and sorting with wavelets and superparamagnetic
clustering. Neural computation 16, 1661–1687 (2004).
23. Chase, S. M. & Young, E. D. First-spie latency information in single neurons increases when referenced to population onset.
Proceedings of the National Academy of Sciences 104, 5175–5180 (2007).
24. onerding, W. S. et al. Encapsulated cell device approach for combined electrical stimulation and neurotrophic treatment of the deaf
cochlea. Hearing Research 350, 110–121 (2017).
25. Lorente De, N. . & Harris, A. S. Experimental studies in hearing. e Laryngoscope 43, 315–326, https://doi.org/10.1288/00005537-
193304000-00013 (1933).
26. Svirsis, G., ota, V., Sanes, D. H. & inzel, J. Enhancement of Signal-to-Noise atio and Phase Locing for Small Inputs by a Low-
reshold Outward Current in Auditory Neurons. e Journal of Neuroscience 22, 11019–11025 (2002).
27. Phillips, D. P., Semple, M. N. & itzes, L. M. Factors shaping the tone level sensitivity of single neurons in posterior eld of cat
auditory cortex. Journal of Neurophysiology 73, 674–686 (1995).
28. Wallace, M. N. & Palmer, A. . Laminar dierences in the response properties of cells in the primary auditory cortex. Experimental
Brain Research 184, 179–191, https://doi.org/10.1007/s00221-007-1092-z (2008).
29. Nelson, M. J. & Pouget, P. Do electrode properties create a problem in interpreting local eld potential recordings? Journal of
Neurophysiology 103, 2315–2317, https://doi.org/10.1152/jn.00157.2010 (2010).
30. edies, H., Sieben, U. & Creutzfeldt, O. D. Functional subdivisions in the auditory cortex of the guinea pig. Journal of Comparative
Neurology 282, 473–488 (1988).
31. Schreiner, C. E., Froeme, . C. & Atencio, C. A. In e auditory cortex (eds J. Winer & C. Schreiner) 275–308 (Springer, 2011).
32. Hellweg, F. C., och, . & Vollrath, M. epresentation of the cochlea in the neocortex of guinea pigs. Experimental Brain Research
29, 467–474, https://doi.org/10.1007/bf00236184 (1977).
33. Wallace, M. N., utowsi, . G. & Palmer, A. . Identication and localisation of auditory areas in guinea pig cortex. Experimental
Brain Research 132, 445–456, https://doi.org/10.1007/s002210000362 (2000).
34. Hener, ., Hener, H. & Masterton, B. Behavioral measurements of absolute and frequency-dierence thresholds in guinea pig.
e Journal of the Acoustical Society of America 49, 1888–1895 (1971).
35. Da llos, P. e auditory periphery biophysics and physiology. (Academic Press, 1973).
36. Evans, E. F. e frequency response and other properties of single bres in the guinea-pig cochlear nerve. e Journal of Physiology
226, 263–287, https://doi.org/10.1113/jphysiol.1972.sp009984 (1972).
37. Norena, A. & Eggermont, J. J. Comparison between local eld potentials and unit cluster activity in primary auditory cortex and
anterior auditory eld in the cat. Hearing Research 166, 202–213, https://doi.org/10.1016/S0378-5955(02)00329-5 (2002).
38. Xing, D., Yeh, C.-I. & Shapley, . M. Spatial Spread of the Local Field Potential and its Laminar Variation in Visual Cortex. e
Journal of Neuroscience 29, 11540–11549, https://doi.org/10.1523/jneurosci.2573-09.2009 (2009).
39. Goodwin, L. D. & Leech, N. L. Understanding Correlation: Factors at Aect the Size of r. e Journal of Experimental Education
74, 249–266, https://doi.org/10.3200/jexe.74.3.249-266 (2006).
40. Pettersen, . H., Lindén, H., Dale, A. M. & Einevoll, G. T. In Handbook of neural activity measurement Vol. 1 (eds omain Brette &
Alain Destexhe) 92–135 (Cambridge University Press, 2012).
41. Csicsvari, J. et al. Massively parallel recording of unit and local field potentials with silicon-based electrodes. Journal of
Neurophysiology 90, 1314–1323 (2003).
42. Bereshpolova, Y., Amitai, Y., Gusev, A. G., Stoelzel, C. . & Swadlow, H. A. Dendritic Bacpropagation and the State of the Awae
Neocortex. e Journal of Neuroscience 27, 9392–9399, https://doi.org/10.1523/jneurosci.2218-07.2007 (2007).
43. Voigt, M. B., Huba, P. & ral, A. Intracortical microstimulation dierentially activates cortical layers based on stimulation depth.
Brain Stimulation 10, 684–694, https://doi.org/10.1016/j.brs.2017.02.009 (2017).
44. Im, C. & Seo, J.-M. A review of electrodes for the electrical brain signal recording. Biomedical Engineering Let ters 6, 104–112, https://
doi.org/10.1007/s13534-016-0235-1 (2016).
45. Eggermont, J. J. & Smith, G. M. Synchrony between single-unit activity and local eld potentials in relation to periodicity coding in
primary auditory cortex. Journal of Neurophysiology 73, 227–245 (1995).
46. Abrams, D. A., Nicol, T., White-Schwoch, T., Zecer, S. & raus, N. Population responses in primary auditory cortex simultaneously
represent the temporal envelope and periodicity features in natural speech. Hearing Research 348, 31–43, https://doi.org/10.1016/j.
heares.2017.02.010 (2017).
Acknowledgements
We thank P. Hubka for support with data analysis.
Author Contributions
W.S.K. and U.P.F. designed the ECoG grid. W.S.K., A.K. and P.B. designed the study. W.S.K. and P.B. performed
the experiments and analyzed the data. W.S.K. performed the statistical analysis and P.B. prepared the gures.
W.S.K. draed the manuscript. All authors discussed the data analysis, reviewed and edited the manuscript.
Additional Information
Supplementary information accompanies this paper at https://doi.org/10.1038/s41598-018-22051-z.
Competing Interests: At the time of the ECoG development, U.P.F. was employed by Blackrock Microsystems
Europe. e authors declare no further conict of interest. is project was supported by the German Federal
Ministry of Education and Research KMU-Innovativ: Medizintechnik FKZ 13GW0050B and by the Deutsche
Forschungsgemeinscha (DFG) EXC 1077/1 Hearing4all. Other than providing nancial support, the funding
sources had no part in the study.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
14
ScIEnTIFIc RePoRtS | (2018) 8:3825 | DOI:10.1038/s41598-018-22051-z
Publisher's note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional aliations.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International
License, which permits use, sharing, adaptation, distribution and reproduction in any medium or
format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Cre-
ative Commons license, and indicate if changes were made. e images or other third party material in this
article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the
material. If material is not included in the article’s Creative Commons license and your intended use is not per-
mitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the
copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.
© e Author(s) 2018
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... The μECoG device with high-density multi-channel electrodes and fabricated by semiconductor technology enables the simultaneous recording of local field potentials (LFPs) from multiple regions in cerebral cortex with high spatiotemporal resolution [15][16][17]. The sheet-shaped μECoG device which has been developed using parylene C to reduce sheet thickness, has improved electrode adhesiveness to the curved surface of the brain [6]. ...
... Indeed, the insertable µECoG was relatively thicker than the ECoG devices reported previously [6], and therefore, it was possible to move the electrodes to the target area in the narrow space between skull and cortex. Furthermore, the recording data obtained by the ECoG device were comparable to those of devices developed previously [16,33]. This indicates that optimization of thickness and stiffness is important for the development of adequate devices. ...
Article
Full-text available
Characterization of inter-regional interactions in brain is essential for understanding the mechanism relevant to normal brain function and neurological disease. The recently developed flexible micro (μ)-electrocorticography (μECoG) device is one prominent method used to examine large-scale cortical activity across multiple regions. The sheet-shaped μECoG electrodes arrays can be placed on a relatively wide area of cortical surface beneath the skull by inserting the device into the space between skull and brain. Although rats and mice are useful tools for neuroscience, current μECoG recording methods in these animals are limited to the parietal region of cerebral cortex. Recording cortical activity from the temporal region of cortex in mice has proven difficult because of surgical barriers created by the skull and surrounding temporalis muscle anatomy. Here, we developed a sheet-shaped 64-channel μECoG device that allows access to the mouse temporal cortex, and we determined the factor determining the appropriate bending stiffness for the μECoG electrode array. We also established a surgical technique to implant the electrode arrays into the epidural space over a wide area of cerebral cortex covering from the barrel field to olfactory (piriform) cortex, which is the deepest region of the cerebral cortex. Using histology and computed tomography (CT) images, we confirmed that the tip of the μECoG device reached to the most ventral part of cerebral cortex without causing noticeable damage to the brain surface. Moreover, the device simultaneously recorded somatosensory and odor stimulus-evoked neural activity from dorsal and ventral parts of cerebral cortex in awake and anesthetized mice. These data indicate that our μECoG device and surgical techniques enable the recording of large-scale cortical activity from the parietal to temporal cortex in mice, including somatosensory and olfactory cortices. This system will provide more opportunities for the investigation of physiological functions from wider areas of the mouse cerebral cortex than those currently available with existing ECoG techniques. Supplementary Information The online version contains supplementary material available at 10.1186/s13041-023-01019-9.
... This array has already been shown to reliably record neuronal activity of the auditory cortex under anesthesia in a final setting. 46 Using this ECoG surface grid array, we analyzed theta (4)(5)(6)(7)(8), alpha (8)(9)(10)(11)(12), beta (12)(13)(14)(15)(16)(17)(18)(19)(20)(21)(22)(23)(24)(25)(26)(27)(28)(29)(30) Hz, also with respect to low (12)(13)(14)(15)(16)(17)(18)(19)(20), high (20-30 Hz)), and gamma (30-100 Hz) frequency band dynamics in the MCtx and SMCtx cortex of freemoving unilateral 6-OHDA lesion rats before, during, and after STN DBS. ...
... In this acute setting, the recording quality of LFP and spatially selective multiunit activity was similar to that of conventional penetrating multielectrode arrays. 46 In this study, we showed that six days after implantation in awake, free-moving rats, recording of high-quality LFP also was possible in high quality from all channels. ...
Article
Objectives: Enhanced beta oscillations in cortical-basal ganglia (BG) thalamic circuitries have been linked to clinical symptoms of Parkinson's disease. Deep brain stimulation (DBS) of the subthalamic nucleus (STN) reduces beta band activity in BG regions, whereas little is known about activity in cortical regions. In this study, we investigated the effect of STN DBS on the spectral power of oscillatory activity in the motor cortex (MCtx) and sensorimotor cortex (SMCtx) by recording via an electrocorticogram (ECoG) array in free-moving 6-hydroxydopamine (6-OHDA) lesioned rats and sham-lesioned controls. Materials and methods: Male Sprague-Dawley rats (250-350 g) were injected either with 6-OHDA or with saline in the right medial forebrain bundle, under general anesthesia. A stimulation electrode was then implanted in the ipsilateral STN, and an ECoG array was placed subdurally above the MCtx and SMCtx areas. Six days after the second surgery, the free-moving rats were individually recorded in three conditions: 1) basal activity, 2) during STN DBS, and 3) directly after STN DBS. Results: In 6-OHDA-lesioned rats (N = 8), the relative power of theta band activity was reduced, whereas activity of broad-range beta band (12-30 Hz) along with two different subbeta bands, that is, low (12-30 Hz) and high (20-30 Hz) beta band and gamma band, was higher in MCtx and SMCtx than in sham-lesioned controls (N = 7). This was, to some extent, reverted toward control level by STN DBS during and after stimulation. No major differences were found between contacts of the electrode grid or between MCtx and SMCtx. Conclusion: Loss of nigrostriatal dopamine leads to abnormal oscillatory activity in both MCtx and SMCtx, which is compensated by STN stimulation, suggesting that parkinsonism-related oscillations in the cortex and BG are linked through their anatomic connections.
... Intracortical electrodes are inserted in the cerebral cortex to record electrical signals. Any invasive electrodes have high spatial and temporal resolutions with the drawback of neuronal damage [15,16]. Electroencephalography (EEG) Hans Berger recorded his first EEG in 1924. ...
... However, invasive methods also come with inherent challenges like: socio-ethical issues and unavailability of a diverse set of subjects, etc. Nevertheless, the steady advancements in innovative methods and designs like thin-film surface electrode array [54] and wireless inductive link powered ECoG data acquisition system [55] may lead to ubiquitous acceptability and implementation of invasive neural analysis methods. ...
Article
Full-text available
Neurointerface or Brain–Machine Interface is a system of devices that utilizes the electrical nature of neural signals to establish communication with the nervous system. This literature survey paper tries to provide a brief overview of various methods and devices that helped reach the advancement of neural interfaces of recent times. The work focuses on collating published research in a variety of fields pertaining to neuro-mechanics, neural rehabilitation, study of neuronal activities, and development of Brain–Computer Interfaces. Previous studies have been categorized into segments describing the practices followed for detecting brainwaves, collecting and interpreting neuronal signals, stimulating brain regions for limb control, audio–visual and tactile percepts, and algorithms. Materials and actuators were used to achieve control of Man–Machine Interfaces through modulating neuronal activities. The article also discusses the current inclination toward invasive modalities and the possibility of integrating Artificial Intelligence with non-invasive technologies to move toward a new age of Brain–Computer Interfaces.
... However, these studies failed 1 to provide clues from the surface activity of the brain because they performed 3D recordings with a multi-channel electrode array in a limited space such as the 'deep brain'. Other efforts have been made to prove the correlation between the surface and deep brain, by implanting ECoG array with inserting neural probe individually 27,28 . However, this recording system is severely restricted in temporal resolution because the time deviation of the electrophysiological signal from the respective devices occurs. ...
Article
Full-text available
Challenges in the understanding of three-dimensional (3D) brain networks by simultaneously recording both surface and intracortical areas of brain signals remain due to the difficulties of constructing mechanical design and spatial limitations of the implanted sites. Here, we present a foldable and flexible 3D neural prosthetic that facilitates the 3D mapping of complex neural circuits with high spatiotemporal dynamics from the intracortical to cortical region. This device is the tool to map the 3D neural transmission through sophisticatedly designed four flexible penetrating shanks and surface electrode arrays in one integrated system. We demonstrate the potential possibilities of identifying correlations of neural activities from the intracortical area to cortical regions through continuous monitoring of electrophysiological signals. We also exploited the structural properties of the device to record synchronized signals of single spikes evoked by unidirectional total whisker stimulation. This platform offers opportunities to clarify unpredictable 3D neural pathways and provides a next-generation neural interface.
Article
Full-text available
Planar microelectrode arrays (MEAs) for – in vitro or in vivo – neuronal signal recordings lack the spatial resolution and sufficient signal‐to‐noise ratio (SNR) required for a detailed understanding of neural network function and synaptic plasticity. To overcome these limitations, a highly customizable three‐dimensional (3D) printing process is used in combination with thin film technology and a self‐aligned template‐assisted electrochemical deposition process to fabricate 3D‐printed‐based MEAs on stiff or flexible substrates. Devices with design flexibility and physical robustness are shown for recording neural activity in different in vitro and in vivo applications, achieving high‐aspect ratio 3D microelectrodes of up to 33:1. Here, MEAs successfully record neural activity in 3D neuronal cultures, retinal explants, and the cortex of living mice, thereby demonstrating the versatility of the 3D MEA while maintaining high‐quality neural recordings. Customizable 3D MEAs provide unique opportunities to study neural activity under regular or various pathological conditions, both in vitro and in vivo, and contribute to the development of drug screening and neuromodulation systems that can accurately monitor the activity of large neural networks over time.
Article
Full-text available
The combination of electrophysiology and neuroimaging methods allows the simultaneous measurement of electrical activity signals with calcium dynamics from single neurons to neuronal networks across distinct brain regions in vivo. While traditional electrophysiological techniques are limited by photo-induced artefacts and optical occlusion for neuroimaging, different types of transparent neural implants have been proposed to resolve these issues. However, reproducing proposed solutions is often challenging and it remains unclear which approach offers the best properties for long-term chronic multimodal recordings. We therefore created a streamlined fabrication process to produce, and directly compare, two types of transparent surface micro-electrocorticography (μECoG) implants: nano-mesh gold structures (m-μECoGs) versus a combination of solid gold interconnects and PEDOT:PSS-based electrodes (pp-μECoGs). Both implants allowed simultaneous multimodal recordings but pp-μECoGs offered the best overall electrical, electrochemical, and optical properties with negligible photo-induced artefacts to light wavelengths of interest. Showing functional chronic stability for up to four months, pp-μECoGs also allowed the simultaneous functional mapping of electrical and calcium neural signals upon visual and tactile stimuli during widefield imaging. Moreover, recordings during two-photon imaging showed no visible signal attenuation and enabled the correlation of network dynamics across brain regions to individual neurons located directly below the transparent electrical contacts.
Chapter
Human intracranial recordings are, with few exceptions, typically confined to recording of local field potentials. Spike detection has been limited to intracortical measurements with microelectrodes or microwires. Here, we put a spotlight on an emerging class of subdural surface arrays densely packed with micro- to mesoscale electrodes. These now combine the benefits of low invasiveness, large coverage and high signal quality with the option of recording spikes in addition to spatially resolved brain waves. We discuss recent work where features of surface spiking activity have been compared to action potential recordings from intracortical electrodes, showing that substantial information about the underlying cortical columns can be retrieved form the epicortical recordings. These novel approaches have potential for large-scale network studies contributing to a deeper understanding of brain function and hold promise for refinement of clinical diagnostic or interventional procedures.
Article
Unlabelled: Flexible and stretchable neural electrodes are promising tools for high-fidelity interfacing with soft and curvilinear brain surface. Here, we describe a flexible and stretchable neural electrode array that consists of polyacrylonitrile (PAN) nanofiber network reinforced gold (Au) film electrodes. Under stretching, the interweaving PAN nanofibers effectively terminate the formation of propagating cracks in the Au films and thus enable the formation of a dynamically stable electrode-tissue interface. Moreover, the PAN nanofibers increase the surface roughness and active surface areas of the Au electrodes, leading to reduced electrochemical impedance and improved signal-to-noise ratio. As a result, PAN nanofiber network reinforced Au electrode arrays can allow for reliable in vivo multichannel recording of epileptiform activities in rats. Supplementary information: The online version contains supplementary material available at 10.1007/s13534-022-00257-5.
Article
Full-text available
The intracerebral local field potential (LFP) is a measure of brain activity that reflects the highly dynamic flow of information across neural networks. This is a composite signal that receives contributions from multiple neural sources, yet interpreting its nature and significance may be hindered by several confounding factors and technical limitations. By and large, the main factor defining the amplitude of LFPs is the geometry of the current sources, over and above the degree of synchronization or the properties of the media. As such, similar levels of activity may result in potentials that differ in several orders of magnitude in different populations. The geometry of these sources has been experimentally inaccessible until intracerebral high density recordings enabled the co-activating sources to be revealed. Without this information, it has proven difficult to interpret a century's worth of recordings that used temporal cues alone, such as event or spike related potentials and frequency bands. Meanwhile, a collection of biophysically ill-founded concepts have been considered legitimate, which can now be corrected in the light of recent advances. The relationship of LFPs to their sources is often counterintuitive. For instance, most LFP activity is not local but remote, it may be larger further from rather than close to the source, the polarity does not define its excitatory or inhibitory nature, and the amplitude may increase when source's activity is reduced. As technological developments foster the use of LFPs, the time is now ripe to raise awareness of the need to take into account spatial aspects of these signals and of the errors derived from neglecting to do so.
Article
Full-text available
Cortical development extensively depends on sensory experience. Effects of congenital monaural and binaural deafness on cortical aural dominance and representation of binaural cues were investigated in the present study. We used an animal model that precisely mimics the clinical scenario of unilateral cochlear implantation in an individual with single-sided congenital deafness. Multiunit responses in cortical field A1 to cochlear implant stimulation were studied in normal-hearing cats, bilaterally congenitally deaf cats (CDCs), and unilaterally deaf cats (uCDCs). Binaural deafness reduced cortical responsiveness and decreased response thresholds and dynamic range. In contrast to CDCs, in uCDCs, cortical responsiveness was not reduced, but hemispheric-specific reorganization of aural dominance and binaural interactions were observed. Deafness led to a substantial drop in binaural facilitation in CDCs and uCDCs, demonstrating the inevitable role of experience for a binaural benefit. Sensitivity to interaural time differences was more reduced in uCDCs than in CDCs, particularly at the hemisphere ipsilateral to the hearing ear. Compared with binaural deafness, unilateral hearing prevented nonspecific reduction in cortical responsiveness, but extensively reorganized aural dominance and binaural responses. The deaf ear remained coupled with the cortex in uCDCs, demonstrating a significant difference to deprivation amblyopia in the visual system.
Article
Full-text available
Electrocorticography grids have been used to study and diagnose neural pathophysiology for over 50years, and recently have been used for various neural prosthetic applications. Here we provide evidence that micro-scale electrodes are better suited for studying cortical pathology and function, and for implementing neural prostheses. This work compares dynamics in space, time, and frequency of cortical field potentials recorded by three types of electrodes: electrocorticographic (ECoG) electrodes, non-penetrating micro-ECoG (μECoG) electrodes that use microelectrodes and have tighter interelectrode spacing; and penetrating microelectrodes (MEA) that penetrate the cortex to record single- or multiunit activity (SUA or MUA) and local field potentials (LFP). While the finest spatial scales are found in LFPs recorded intracortically, we found that LFP recorded from μECoG electrodes demonstrate scales of linear similarity (i.e., correlation, coherence, and phase) closer to the intracortical electrodes than the clinical ECoG electrodes. We conclude that LFPs can be recorded intracortically and epicortically at finer scales than clinical ECoG electrodes are capable of capturing. Recorded with appropriately scaled electrodes and grids, field potentials expose a more detailed representation of cortical network activity, enabling advanced analyses of cortical pathology and demanding applications such as brain-computer interfaces. Copyright © 2015 International Federation of Clinical Neurophysiology. Published by Elsevier Ireland Ltd. All rights reserved.
Article
Full-text available
Absolute and frequency‐difference thresholds were determined by the conditioned‐suppression technique. The results show that the average frequency range of audibility at +50 dB sound‐pressure level extends from 86 Hz to 46.5 kHz, with a best frequency near 8 kHz. Individual differences in sensitivity are related to body weight and, probably, age. The average frequency‐difference limen is 3.5% from 125 Hz to 42 kHz. Compared to other mammals, the auditory capacities of guinea pig are within one standard deviation of the mammalian mean on each of six dimensions: high‐frequency and low‐frequency cutoff, lowest intensity, best frequency, area of the audible field, and frequency discrimination.
Article
Profound hearing impairment can be overcome by electrical stimulation (ES) of spiral ganglion neurons (SGNs) via a cochlear implant (CI). Thus, SGN survival is critical for CI efficacy. Application of glial cell line-derived neurotrophic factor (GDNF) has been shown to reduce SGN degeneration following deafness. We tested a novel method for local, continuous GDNF-delivery in combination with ES via a CI. The encapsulated cell (EC) device contained a human ARPE-19 cell-line, genetically engineered for secretion of GDNF. In vitro, GDNF delivery was stable during ES delivered via a CI. In the chronic in vivo part, cats were systemically deafened and unilaterally implanted into the scala tympani with a CI and an EC device, which they wore for six months. The implantation of control devices (same cell-line not producing GDNF) had no negative effect on SGN survival. GDNF application without ES led to an unexpected reduction in SGN survival, however, the combination of GDNF with initial, short-term ES resulted in a significant protection of SGNs. A tight fibrous tissue formation in the scala tympani of the GDNF-only group is thought to be responsible for the increased SGN degeneration, due to mechanisms related to an aggravated foreign body response. Furthermore, the fibrotic encapsulation of the EC device led to cell death or cessation of GDNF release within the EC device during the six months in vivo. In both in vitro and in vivo, fibrosis was reduced by CI stimulation, enabling the neuroprotective effect of the combined treatment. Thus, fibrous tissue growth limits treatment possibilities with an EC device. For a stable and successful long-term neurotrophic treatment of the SGN via EC devices in human CI users, it would be necessary to make changes in the treatment approach (provision of anti-inflammatories), the EC device surface (reduced cell adhesion) and the ES (initiation prior to fibrosis formation).
Article
Brain-machine interfaces (BMIs) combine methods, approaches, and concepts derived from neurophysiology, computer science, and engineering in an effort to establish real-time bidirectional links between living brains and artificial actuators. Although theoretical propositions and some proof of concept experiments on directly linking the brains with machines date back to the early 1960s, BMI research only took off in earnest at the end of the 1990s, when this approach became intimately linked to new neurophysiological methods for sampling large-scale brain activity. The classic goals of BMIs are 1) to unveil and utilize principles of operation and plastic properties of the distributed and dynamic circuits of the brain and 2) to create new therapies to restore mobility and sensations to severely disabled patients. Over the past decade, a wide range of BMI applications have emerged, which considerably expanded these original goals. BMI studies have shown neural control over the movements of robotic and virtual actuators that enact both upper and lower limb functions. Furthermore, BMIs have also incorporated ways to deliver sensory feedback, generated from external actuators, back to the brain. BMI research has been at the forefront of many neurophysiological discoveries, including the demonstration that, through continuous use, artificial tools can be assimilated by the primate brain’s body schema. Work on BMIs has also led to the introduction of novel neurorehabilitation strategies. As a result of these efforts, long-term continuous BMI use has been recently implicated with the induction of partial neurological recovery in spinal cord injury patients.
Article
Background: Intracortical microstimulation is one of the most common techniques to causally interfere with neuronal processing, but neuronal recordings spanning the whole cortical depth during stimulation are exceptionally rare. Objective/hypothesis: Here we combined layer-specific intracortical microstimulation with extracellular recordings on the same shank of a linear multi-electrode array to study the effects of electrical stimulation in different cortical depths on intracortical processing in the auditory cortex in vivo. Methods: Population responses (local field potentials and multi-unit activity) were recorded from the auditory cortex of 8 guinea pigs under ketamine/xylazine anesthesia while single current pulses (charge-balanced, biphasic, square-wave, 0.1-45 μA, 200 μs/phase) were delivered in different cortical depths. Results: The cortical responses differed with a change in the stimulation parameters, with significant factors being the stimulating current (p < 0.0001), stimulation depth (p = 0.03) and the recording depth (p = 0.002) considering the local field potential amplitude. A cross-correlation analysis between responses evoked by intracortical microstimulation and physiological auditory stimuli revealed the closest match when stimulating the middle granular layer (p < 0.05). Conclusion: Intracortical response profiles to low-current intracortical microstimulation were layer specific. The most natural cortical response was achieved by stimulation in the thalamo-recipient layer. These findings contribute to a basis for designing cortical neuroprosthetics.
Article
Speech perception relies on a listener’s ability to simultaneously resolve multiple temporal features in the speech signal. Little is known regarding neural mechanisms that enable the simultaneous coding of concurrent temporal features in the speech signal. Here we show that two categories of temporal features in speech, the low-frequency speech envelope and periodicity cues, are processed by distinct neural mechanisms within the same population of cortical neurons. We measured population activity in primary auditory cortex of anesthetized guinea pig in response to three variants of a naturally produced speech sentence. Results show that the envelope of population responses closely tracks the speech envelope, and this cortical activity more closely reflects wider bandwidths of the speech envelope compared to narrow bands. Additionally, neuronal populations represent the fundamental frequency of speech robustly with phase-locked responses. Importantly, these two temporal features of speech are simultaneously observed within neuronal ensembles in auditory cortex in response to clear, conversation, and compressed speech exemplars. Results show that auditory cortex is adept at simultaneously resolving multiple temporal features in extended speech sentences using discrete coding mechanisms.
Article
Brain is complex organ composed of numerous glial cells and neurons to convey information using chemical and electrical signals. Neural interface technology using the electrical brain signals has attracted great attention for the clinical and experimental applications. Electrode as the neural interface is the most important part in stimulating neural cells or recording neural activities. In this paper, we provide an overview of electrodes for recording the electrical brain signal. The noninvasive electrodes are primarily used to capture electroencephalogram (EEG) from outside the skull while the implantable electrodes are employed to measure electrocorticogram (ECoG), local field potential (LFP) or spike activity. Recent progress in microfabrication technology enables the development of on-board electrode that combines the entire signal processing including amplification, filtering, and digitization. This will contribute to diagnostic and therapeutic application of the neural interface for restoring physical, psychological and social functions by improving motor, sensory or cognitive abilities.
Article
Background: Current source density analysis of recordings from penetrating electrode arrays has traditionally been used to examine the layer-specific cortical activation and plastic changes associated with changed afferent input. We report on a related analysis, the second spatial derivative (SSD) of surface local field potentials (LFPs) recorded using custom designed thin-film polyimide substrate arrays. Results: SSD analysis of tone-evoked LFPs generated from the auditory cortex under the recording array demonstrated a stereotypical single local minimum, often flanked by maxima on both the caudal and rostral sides. In contrast, tone-pips at frequencies not represented in the region under the array, but known (on the basis of normal tonotopic organization) to be represented caudal to the recording array, had a more complex pattern of many sources and sinks. Comparison with existing methods: Compared to traditional analysis of LFPs, SSD analysis produced a tonotopic map that was more similar to that obtained with multi-unit recordings in a normal-hearing animal. Additionally, the statistically significant decrease in the number of acoustically responsive cortical locations in partially deafened cats following 6 months of cochlear implant use compared to unstimulated cases observed with multi-unit data (p=0.04) was also observed with SSD analysis (p=0.02), but was not apparent using traditional analysis of LFPs (p=0.6). Conclusions: SSD analysis of surface LFPs from the thin-film array provides a rapid and robust method for examining the spatial distribution of cortical activity with improved spatial resolution compared to more traditional LFP recordings.