Download Corina Wirth and Hans

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Cortical cooling wikipedia , lookup

Endocannabinoid system wikipedia , lookup

Electrophysiology wikipedia , lookup

Neurotransmitter wikipedia , lookup

Stimulus (physiology) wikipedia , lookup

Molecular neuroscience wikipedia , lookup

Activity-dependent plasticity wikipedia , lookup

Executive functions wikipedia , lookup

Metastability in the brain wikipedia , lookup

Environmental enrichment wikipedia , lookup

Cognitive neuroscience of music wikipedia , lookup

Neural modeling fields wikipedia , lookup

Time perception wikipedia , lookup

Microneurography wikipedia , lookup

Neuropsychopharmacology wikipedia , lookup

Neuroeconomics wikipedia , lookup

Apical dendrite wikipedia , lookup

Single-unit recording wikipedia , lookup

Anatomy of the cerebellum wikipedia , lookup

Neural correlates of consciousness wikipedia , lookup

Transcranial direct-current stimulation wikipedia , lookup

Premovement neuronal activity wikipedia , lookup

Eyeblink conditioning wikipedia , lookup

Optogenetics wikipedia , lookup

Spike-and-wave wikipedia , lookup

Evoked potential wikipedia , lookup

Feature detection (nervous system) wikipedia , lookup

Channelrhodopsin wikipedia , lookup

Cerebral cortex wikipedia , lookup

Neurostimulation wikipedia , lookup

Multielectrode array wikipedia , lookup

Synaptic gating wikipedia , lookup

Transcript
J Neurophysiol 91: 1635–1647, 2004.
First published November 19, 2003; 10.1152/jn.00950.2003.
Spatiotemporal Evolution of Excitation and Inhibition in the Rat Barrel
Cortex Investigated With Multielectrode Arrays
Corina Wirth and Hans-R. Lüscher
Institute of Physiology, University of Bern, CH-3012 Bern, Switzerland
Submitted 2 October 2003; accepted in final form 13 November 2003
Since the discovery of the “barrel” map in the rodent somatosensory cortex by Woolsey and Van der Loos (1970), the
barrels were regarded as a manifestation of the cortical columns in layer 4 (L4). It turned out that this structure is a
well-suited model for investigating sensory information processing in the cerebral cortex. L4 is the main input layer for
sensory information relayed through the thalamus and is therefore considered as the first stage in cortical signal processing.
Barrels receive input from the ventral posterior medial (VPM)
nucleus (Hirsch 1995; Jensen and Killackey 1987; for a review,
see Diamond 1995) and are thought to be essential to extract
relevant thalamic signals and to amplify (Feldmeyer et al.
1999) or even to damp thalamic input (Pinto et al. 2003).
Axonal and dendritic arbors of barrel neurons are confined to
barrel borders (Feldmeyer et al. 1999; Lübke et al. 2000;
Petersen and Sakmann 2000), suggesting that barrels work as
functionally independent excitatory neuronal networks (Gold-
reich et al. 1999; Laaris and Keller 2002; Petersen and Sakmann 2001). L4 cells also target pyramidal cells in L2/3
(Armstrong-James et al. 1992; Feldmeyer et al. 2002) whose
axons arborize in L2/3 and in L5 (Gottlieb and Keller 1997).
Signal processing in L2/3 is no longer confined to a barrel
column but spread in vertical and horizontal directions, in
particular to L5 (Feldmeyer et al. 2002; Laaris et al. 2000;
Lübke et al. 2000; Petersen and Sakmann 2001).
Excitatory thalamic input induces a rapid and strong feedforward inhibition, limiting spread of excitation both spatially
and temporally. Some inhibitory L4 interneurons are excited
directly by the thalamocortical afferents (Bruno and Simons
2002; Galarreta and Hestrin 2002; Porter et al. 2001; Staiger et
al. 1996) and probably limit the duration of the excitation
generated by thalamocortical input. In addition, L4 interneurons receive input from local excitatory neurons and inhibit
excitatory L4 neurons, participating in feedback inhibition
(Amitai et al. 2002; Beierlein et al. 2000; Gibson et al. 1999).
Interneurons are also interconnected by GABAergic synapses
and gap junctions and, in addition, form autapses (Bacci et al.
2003; Gibson et al. 1999). Porter et al. (2001) found a preponderance of L4 interneurons directed to L2/3 suggesting that the
flow of inhibition in the barrel cortex after thalamocortical
input could be toward the upper layers, probably following
excitation along its intracortical course being inseparably
linked to it. However, the precise spatiotemporal behavior of
the inhibitory network in L4 and L2/3 has never been described
in detail.
Another controversial subject is the role of infragranular
layers. L5/6 is generally considered as the main output layer,
processing information arriving mainly from L2/3, but direct
input in L6 after thalamic stimulation has been observed as
well (Armstrong-James et al. 1992; Beierlein and Connors
2002; Laaris et al. 2000). Interestingly, corticothalamic cells in
L6 also send collaterals to L4 (Staiger et al. 1996; Zhang and
Deschenes 1997), and axons of L4 neurons were observed to
end in L5/6 (Laaris and Keller 2002; Lübke et al. 2000;
Petersen and Sakmann 2000). These anatomical observations
expand the generally postulated simple “thalamus-L4-L2/3L5/6” pathway.
To gain more insight in the spatiotemporal spread of activity
in rat barrel cortex, we used multielectrode arrays with 60
pyramidal shaped electrodes permitting simultaneous multisite
recordings in all layers of rat barrel cortex in vitro. Stimulation
of individual barrels confirmed the hypothesis that barrels
operate as individual entities. Excitation spreading from L4 to
Address for reprint requests and other correspondence: H.-R. Lüscher,
Institute of Physiology, University of Bern, Bühlplatz 5, CH-3012 Bern,
Switzerland (E-mail: [email protected]).
The costs of publication of this article were defrayed in part by the payment
of page charges. The article must therefore be hereby marked ‘‘advertisement’’
in accordance with 18 U.S.C. Section 1734 solely to indicate this fact.
INTRODUCTION
www.jn.org
0022-3077/04 $5.00 Copyright © 2004 The American Physiological Society
1635
Downloaded from http://jn.physiology.org/ by 10.220.33.5 on June 18, 2017
Wirth, Corina and Hans-R. Lüscher. Spatiotemporal evolution of
excitation and inhibition in the rat barrel cortex investigated with
multielectrode arrays. J Neurophysiol 91: 1635–1647, 2004. First
published November 19, 2003; 10.1152/jn.00950.2003. We investigated the spatiotemporal evolution of activity in the rat barrel cortex
using multielectrode arrays (MEAs). In acute brain slices, field potentials were recorded simultaneously from 60 electrodes with high
spatial and temporal resolution. This new technique allowed us to map
functionally discrete barrels and to observe the interplay between the
excitatory and inhibitory network. The local field potentials (LFPs)
were elicited by focal electrical stimulation in layer 4 (L4). Excitation
recorded in a single barrel was first confined to the stimulated barrel
and subsequently spread in a columnar manner to layer 2/3 (L2/3).
This excitation in L4 and lower L2/3 was followed by inhibition
curtailing excitation to a short period lasting only ⬃2 ms. In the
uppermost layer, a long-lasting (⬃10 ms), laterally spreading band of
excitation remained active. Blockade of GABAA-receptors resulted in
a long-lasting and diffuse activation of L4 and lower L2/3 and
abolition of activation of the upper L2/3. Thus inhibition not only
shaped the spatial-temporal map of excitation in L4 and lower L2/3
but also resulted indirectly in an excitatory action in the superficial
layers. Stimulation in L6 revealed a feedforward inhibition to L4 and
subsequently an excitatory L6-L4-L6 loop. The complex interplay
between excitation and inhibition opens two spatial windows of
excitation in the infra- and supragranular layers. They may prepare the
L5 pyramidal neuron for associating top-down input from other cortical regions with bottom-up input from the whisker pad to generate
behaviorally relevant output.
1636
C. WIRTH AND H.-R. LÜSCHER
supragranular layers is temporally confined by inhibition, leaving only a brief window of excitation in lower L2/3. A similar
but longer lasting window of excitation in L5/6 is opened by a
specific L5/6-L4-L5/6 loop. In the upper-most layer, excitation
is longer lasting probably due to the excitatory action of
GABAergic input to the distal dendritic region of pyramidal
cells (Gulledge and Stuart 2003). The excitatory action in the
distal dendrite could lower the threshold for backpropagation
induced Ca2⫹ spike firing (Larkum et al. 1999). This spatial
pattern of inhibition and excitation puts the L5 pyramidal
neuron into the position to associate top-down input from other
cortical regions with bottom-up input from the whisker pad to
generate behaviorally relevant output.
METHODS
Twelve- to 20-days-old Wistar rats were decapitated and 350 ␮m
thick thalamocortical slices were prepared in ice-cold artificial cerebrospinal fluid (ACSF) with a vibratome (D.S.K. microslicer DTK1000, Dosaka, Kyoto, Japan) as described by Agmon and Connors
(1991). They were incubated in ACSF at 34° for 1 h after slicing. The
preparation of the slices was done according to federal and institutional guidelines. ACSF contained (in mM) 125 NaCl, 2.5 KCl, 1.25
NaH2PO4, 1 MgCl2, 2 CaCl2, 25 NaHCO3, and 25 glucose and was
bubbled with 95% O2-5% CO2.
Identification of barrels in living slices
In thalamocortical slices, barrels can be recognized even by the
naked eye. With a ⫻2.5 lens, the barrel field is clearly visible as a
band of alternating dark and light patches in L4. We used only slices
with clearly distinguishable barrels. Due to optical interference with
the nontransparent electrodes and leads of the multielectrode array
(MEA) the barrels were often not clearly visible after placing the
slices on the MEA (Fig. 1A). We only kept slices for further experi-
Multielectrode arrays
Arrays of 60 pyramidal shaped electrodes (Ti/Au/TiN) with the
following properties were used (Fig. 1A): electrode diameter, 40 ␮m;
electrode spacing, 140 ␮m; tip height, 42 ␮m; impedance, 400 – 600
k⍀ (Ayanda Biosystems SA, Lausanne, Switzerland). The MEA was
covered with 0.1% polyethylenimine (PEI) dissolved in distilled water. Excess PEI solution was removed after ⬃2 h. The MEA was
rinsed with distilled water and left to dry. This coating procedure
improved the contact between MEA electrodes and the brain slice,
resulting in a larger signal to noise ratio. Prior to recording the slices
were trimmed to fit approximately the size of the MEA (⬃3 ⫻ 3 mm).
This trimming removed most of the thalamic structure from the slice.
After placing the slice on the MEA, we carefully removed all ACSF
between the array and the slice with blotting paper. Afterward we
added a moist ring of blotting paper around the slice and covered the
slice mounted on the array with a plastic cap. We kept the slice for 1–2
min in this humid chamber. This procedure improved the contact
between slice and MEA. Afterward, we added ACSF to the slice until
it was completely submerged and started perfusion at a temperature of
35°C.
Stimulation
We used ACSF-filled double-barrel micropipettes (␪-glass) with tip
diameters of 5–10 ␮m for extracellular stimulation. Constant amplitude voltage pulses (0.5–5 V, 150 ␮s, 0.3 Hz) were delivered with a
modified optically isolated stimulator (Iso Flex, A.M.P.I., Israel)
driven by a timer (Master-8, A.M.P.I, Israel). Stimulus intensity was
1.2–1.5 times threshold in all recordings. Individual barrels were
stimulated by placing the tip of the stimulus-pipette in the center of
the barrels. In a second series of experiments, we stimulated at
individual locations along a horizontal line through L4 in steps of 100
␮m taking the MEA electrodes as landmarks. If other stimulus locations were used, it is mentioned in RESULT.
FIG. 1. Thalamocortical slices and multielectrode array
(MEA) recordings. A: small piece of barrel field cortex mounted
on the MEA. L1 covers the 1st MEA row. Row 13 of the MEA
is located deep in layer 6 at the border to the white matter. The
barrels are barely visible as bright clusters in rows 5–7 of the
MEA. The leads connecting the individual electrodes to the
connectors of the preamplifier can clearly be seen. B: stimulus
induced (3V, 0.3 Hz, 150 ␮s) local field potentials (LFPs) recorded with the MEA electrodes (average over 20 sweeps). The
stimulus artifact indicates the timing of the stimulus, which was
delivered via a theta-glass micropipette filled with ACSF at the
location indicated by the asterisk. Three damaged electrodes in
the right lower corner of the MEA were switched off. C: spatial
activity map of the postsynaptic signals of the recording shown in
B, 3 ms after stimulation (see METHODS). Black dot: stimulation
location. D: enlarged recordings from the single electrode indicated in B (average over 20 sweeps). To extract the postsynaptic
response from the LFP (blue trace), we subtracted the presynaptic
component recorded in the presence of 200 ␮M Cd2⫹ and 100
␮M Ni2⫹ (green trace) from the original trace (red). This procedure eliminated the fiber volleys and terminal potentials leaving
the postsynaptically induced synaptic local field potentials
(sLFPs) and was applied to all recordings prior to the analysis
shown in this paper.
J Neurophysiol • VOL
91 • APRIL 2004 •
www.jn.org
Downloaded from http://jn.physiology.org/ by 10.220.33.5 on June 18, 2017
Slice preparation
ments if at least one barrel was visually recognizable after mounting
on the MEA.
SPATIOTEMPORAL EVOLUTION OF ACTIVITY IN BARREL CORTEX
Data acquisition
A 60-channel amplifier was used with a band-pass filter set between
10 Hz and 3 kHz, 1200⫻ amplification (B-MEA-1060, Multi Channel
Systems, Reutlingen, Germany), and data were acquired by the PCbased data software MC Rack (Multi Channel Systems) at a sampling
rate of 10 kHz.
ms. The center of mass sv(t) at time t within the vertical dimension (in
units of MEA rows) was obtained by
sv共t兲 ⫽
sh共t兲 ⫽
The local field potential (LFP) at each electrode was recorded
against the bath electrode. To improve the signal-to-noise ratio, the
traces were averaged (n ⫽ 15–20). Because spike activity can be
detected with MEA electrodes at distances of ⱕ100 ␮m (Egert et al.
2002), we assume that each electrode records additional weak signals
from neurons located close to neighboring electrodes. Thus MEA
electrodes with a spacing of 140 ␮m cannot be considered as completely independent recording units. As a correction procedure we
routinely subtracted 10% of the voltage recorded by each of the four
neighboring electrodes from the voltage recorded by the center electrode. This procedure improves the signal contrast slightly without
affecting the overall activity pattern. The 60 recordings were displayed in a spatial map corresponding to the electrode location (Fig.
1B). The individual traces are displayed in such a way as to mimic the
polarity of the underlying synaptic potentials identified by pharmacological means (see RESULTS, Fig. 2). The trace consisted of a shortlatency, brief positive signal immediately after the stimulus artifact
that was followed by a broader positive peak. This second positive
peak was followed by a negative peak (Fig. 1D, red trace). At the end
of each experiment 200 ␮M Cd2⫹ and 100 ␮M Ni2⫹ were added to
the superfusion solution and the experiment repeated. By subtracting
the traces obtained under Ni2⫹/Cd2⫹ (Fig. 1D, green trace) from the
traces recorded under control conditions, the short-latency peak almost completely disappeared, suggesting that it represents a “fiber
volley” or axon terminal potentials (Swadlow and Gusev 2000)
evoked by direct stimulation of the axons. We call these reduced
traces synaptic LFPs (sLFPs). The sLFPs were exclusively used for
further analysis (Fig. 1D, blue trace).
To visualize the spatial spread of the signals, we generated colorcoded activity maps representing the instantaneous voltage amplitude
of the sLFP recorded by each MEA electrode using cubic interpolation for amplitude ratios at intermediate positions (Fig. 1C). Red
always coded for a positive signal amplitude, whereas blue coded for
a negativity in the individual sLFP. With a sampling frequency of 10
kHz, such an activity map could be generated every 0.1 ms. To
document the temporal evolution of the sLFP, we produced time
series of activity maps with a time interval of 0.2 or 0.4 ms between
successive pictures. They were obtained by taking the average of the
two or four activity maps were calculated for a time interval of 0.1 ms.
To represent a laminar activity profile along the vertical axes of the
cortical column, we calculated the average of the activity across the
MEA columns covering the cortical column defined by the activated
barrel. These profiles were color coded and displayed as a line scan
with time on the abscissa and space across the cortical thickness as the
ordinate (see Fig. 4A2). For a more detailed analysis of the spatiotemporal evolution of the activity, we calculated the center of mass of
the signal amplitude over all recording electrodes obtained every 0.1
J Neurophysiol • VOL
冘
mij共t兲 䡠 i
(1)
i, j
1
M共t兲
冘
mij共t兲 䡠 j
(2)
i, j
where M(t) ⫽ ¥mij and mij(t) represents the thresholded voltage
i,j
recorded at time t with the MEA electrode at row i and column j. The
summation was performed over all MEA electrodes in the chosen
rows and columns. To analyze the excitatory signals (e.g., Fig. 3C),
the thresholding was performed by subtracting a noise level of 8 ␮V
and considering only the positive values for mij(t). For inhibitory
signals (e.g., Fig. 5B2), a noise level of ⫺8 ␮V was subtracted and
only the negative values were considered for mij(t).
To quantify the voltage distribution around the center of mass, we
calculated the SD for the horizontal dimension given by
␴ h 共t兲 ⫽
冑 冘
1
M共t兲
共共s共t兲 ⫺ j兲2m ij 兲
(3)
ij
RESULTS
We have organized the result section in the following way.
First, a pharmacological analysis of the sLFPs recorded by the
individual MEA electrodes will be presented and the excitatory
and inhibitory signal components will be determined. Second,
evidence that the early excitatory and inhibitory activity remains confined to the barrel borders will be shown. Third, the
basic features of the temporal evolution of the activity maps
after electrical stimulation of individual barrels will be presented. Fourth, the spatiotemporal evolution of the activity will
be dissected into its constituent excitatory and inhibitory components and the spatial and temporal shaping of activity by
inhibition will be analyzed in the different cortical layers.
Finally, the spatiotemporal evolution of the activity will be
presented after stimulation of L6.
Pharmacological characterization of stimulus induced LFPs
In a first set of experiments, we stimulated in L4, the main
input layer to the somatosensory cortex. Individual traces of
the sLFP consisted of a positive peak 3.5 ⫾ 0.8 (SD) ms (n ⫽
43 experiments) after stimulation that was followed in 53% of
the experiments by a negative peak 6.4 ⫾ 2.4 ms after stimulation (Fig. 1D, blue trace; Fig. 2, A–C). The positive component was present around the stimulation location in L4 (mean
peak amplitude in L4: 24.1 ⫾ 20.8 ␮V, n ⫽ 43 experiments)
but was also detected at other locations (mean peak amplitude
at all locations: 14.1 ⫾ 5.2 ␮V, n ⫽ 43 experiments.) The
negative component was primarily present in L4 and lower
L2/3 but never in upper L2/3 (mean peak amplitude at all
locations: ⫺15.6 ⫾ 6.4 ␮V, n ⫽ 23 experiments). The amplitude and temporal relation between the positive and the negative peak was very similar from trial to trial but varied considerably from experiment to experiment. The traces in Fig. 2,
A and C, give examples of two extremes.
The different components of the sLFP were pharmacologically dissected using in a first set of experiments the GABAA-
91 • APRIL 2004 •
www.jn.org
Downloaded from http://jn.physiology.org/ by 10.220.33.5 on June 18, 2017
Data analysis
1
M共t兲
and the center of mass sh(t) at time t within the horizontal dimension
(in units of MEA columns) was obtained by
Pharmacology
Stock solutions of the following drugs were dissolved in ACSF
getting the final concentrations given in the following text: 10 ␮M
bicuculline, 10 ␮M 6-cyano-7-nitroquinoxaline-2,3-dione (CNQX),
50 ␮M D(⫺)-2-amino-5-phosphonopentanoic acid (D-AP5), and 1 ␮M
tetrodotoxin (TTX). Synaptic transmission was blocked by adding 200
␮M Cd2⫹ and 100 ␮M Ni2⫹ to the superfusion solution (all drugs and
chemicals were from Sigma, Merck, Tocris, or Alamone).
1637
1638
C. WIRTH AND H.-R. LÜSCHER
but this observation was not analyzed systematically. The
negative component following the positive signal disappeared
almost completely after the application of CNQX and D-AP5.
The small negative component remaining after the application
of CNQX and D-AP5 disappeared after application of 10 ␮M
bicuculline (Fig. 2C) and must therefore be evoked by direct
stimulation of inhibitory fibers. This analysis suggests that the
positive and negative sLFP components in L4 can be related to
local excitatory and inhibitory postsynaptic potentials. It does,
however, not exclude the possibility that the sLFP is contaminated by passive current flow from distant sources.
Early activation remained spatially confined to the
individual barrel
receptor antagonist bicuculline (10 ␮M, n ⫽ 11). After application of bicuculline, the negative sLFP peak recorded in L4
was completely or largely abolished (Fig. 2A). In the cases
where no negative component was seen, application of bicuculline led to a broadening and increase of the positive peak.
These findings suggest that the negative component of the
sLFP was mainly due to GABAA-receptor activation. Second,
we blocked the AMPA and NMDA receptors by applying 10
␮M CNQX and 50 ␮M D-AP5, respectively (n ⫽ 6). The
positive peak decreased by 21 ⫾ 15% after the application of
D-AP5 and vanished completely after adding CNQX (Fig. 2, B
and C), suggesting that the positive sLFP component was
caused by both AMPA and NMDA receptor activation. We
have observed regional differences in the NMDA/AMPA ratio,
J Neurophysiol • VOL
91 • APRIL 2004 •
www.jn.org
Downloaded from http://jn.physiology.org/ by 10.220.33.5 on June 18, 2017
FIG. 2. Pharmacological analysis of stimulus induced sLFP. A: recordings
from a single MEA electrode under artificial cerebrospinal fluid (ACSF; red
trace) and under 10 ␮M bicuculline (blue trace). The artifact at 5 ms coincided
with the time of stimulation. B: recording from a single MEA electrode under
control conditions (red trace) and with 50 ␮M D-AP5 and 10 ␮M CNQX (blue
trace). C: recording from a single MEA electrode under control condition (red
trace), 50 ␮M D-AP5 (blue trace), additional 10 ␮M CNQX (green trace) and
additional 10 ␮M bicuculline (black trace). The small negative component
remaining in B disappeared after addition of bicuculline.
It has been shown previously by means of voltage-sensitive
dyes that barrels are functionally independent units (Laaris and
Keller 2002; Petersen and Sakmann 2001). To validate the new
recording technique presented in this paper, we performed
similar experiments with MEA recordings as those reported by
Petersen and Sakmann (2001), who used voltage-sensitive
dyes. The stimulation electrode was moved stepwise along a
horizontal line through L4 in steps of 100 ␮m applying at each
location identical stimuli (0.5–5V, 150 ␮s, 0.3 Hz; n ⫽ 11).
The first positive sLFP peaks occurred in L4 and were
limited to small areas corresponding largely to the anatomically defined barrels seen in the microscope image (Fig. 3, A
and B), in agreement with recent findings (Laaris and Keller
2002; Petersen and Sakmann 2001). Some barrels could be
fully activated even when the stimulation electrode was placed
at the edge of a barrel, whereas other barrels could only be
activated partially or not at all (Fig. 3B).
The time from stimulus onset to the peak of the activation of
single barrels varied considerably (Fig. 3B), lasting from 2 to
5 ms after stimulation (3.5 ⫾ 0.8 ms, n ⫽ 43).
An activity profile across the stimulated barrel was calculated by summing up the activity in L4 at the time of maximal
activation. Ten activity profiles from different barrels are superimposed after the peak intensity for each of the 8 experiments was normalized (Fig. 3D). The mean of these 10 activity
profiles was fitted by a Gaussian curve yielding a SD of 95 ␮m.
Two times the SD gives a rough estimate of the average barrel
width of ⬃200 ␮m. To analyze the temporal evolution of the
signals, we calculated the center of mass of the positive response in L4 (see METHODS) and plotted its location against time
after stimulation (Fig. 3C, 9 experiments). This figure illustrates the horizontal confinement of the centers of mass to a
small corridor of ⬃200 ␮m corresponding to the preceding
estimated average barrel width. The center of mass respected
the barrel borders as long as the stimulus location was within
the barrel borders. The cluster-like activation of L4 caused by
stimulation in 100 ␮m-steps along L4 was demonstrated by
calculating the average of the center of mass for the time
interval 1.5–5 ms after stimulation and plotting it along the
MEA rows (Fig. 3E, 5 different experiments, each represented
by a different color). This clearly demonstrates that the centers
of mass are clustered within the anatomically defined barrels
and that the activation respects the individual barrel borders.
SPATIOTEMPORAL EVOLUTION OF ACTIVITY IN BARREL CORTEX
1639
Spatiotemporal evolution of activity after stimulation
of single barrels
Stimulating an individual barrel in its barrel hollow led to a
stereotypical spatiotemporal evolution of the activity. 1.8 ⫾
0.6 ms (n ⫽ 43) after stimulation, excitation was visible in the
stimulated barrel. This short-lasting excitation (on average
1.8 ⫾ 1.2 ms, n ⫽ 43) was replaced in 60% of the experiments
by a slowly increasing inhibitory signal. Before the excitation
in the barrel was replaced by inhibition, excitation spread
within a short time interval from the stimulated barrel to L2/3
above the stimulated barrel in 95% of the experiments (Fig. 4A,
1 and 2). While excitation respected the columnar organization
delineated by the barrel borders in L4 and lower L2/3 immediately above the barrel, it became subsequently broader in
L2/3 (Fig. 4, A1 and B). In the upper L2/3, it spread horizontally and crossed the borders of the cortical column, designing
a mushroom-like activation in L2/3. Inhibition started in the
stimulated barrel 3.4 ⫾ 1.6 ms (n ⫽ 23) after stimulation and
spread from L4 to the lower L2/3 after excitation (see Fig.
4A1). Inhibition in lower L2/3 was observed in all experiments
even if it was not visible in the stimulated barrel, creating often
two distinct regions of excitation, one in the barrel itself, the
other in the upper L2/3.
While excitation of L4 was terminated 3.6 ⫾ 1.8 ms (n ⫽
43) after stimulation, the depolarizing signal in upper L2/3 was
active for as long as 15 ms (9.6 ⫾ 3.0 ms, n ⫽ 30) after
stimulation (Fig. 4A2). The inhibitory signal in lower L2/3 was
visible for 14.7 ⫾ 3.4 ms (n ⫽ 34). The described temporal
evolution of the excitatory and inhibitory components of the
sLFP through the cortical layers can best be followed in Fig.
4A2, where the mean activity across a cortical column is
plotted as a function of time. In addition to the excitation
moving up in to the supragranular layers, a weak excitatory
signal spread to the infragranular layers as well (see also
Fig. 5A1).
J Neurophysiol • VOL
In a small number of experiments (n ⫽ 5, 12%), excitation
in a neighboring barrel could be observed (5.1 ⫾ 1.1 ms) after
stimulation (Fig. 4, B and C). In all these cases, the neighboring
barrel was activated through L2/3, suggesting that signals in
adjacent barrels are primarily evoked by activation of a L4L2/3-L4 pathway and not by direct activation. In two of these
five experiments, a symmetrical activation (although with different delays) of both neighboring barrels could be observed
(not shown).
In 25% of the experiments (n ⫽ 11), excitation propagated
first from the stimulated barrel in L4 to the lower L2/3 and less
than one millisecond later to L5/6 (Fig. 4C). Excitation remained for 8 ⫾ 1.5 ms in L5/6 and did not respect the
columnar organization. In five cases, excitation in L5/6 was
followed by weak but long-lasting negative signals lasting
from 12.2 ⫾ 1.5 to 21.4 ⫾ 4.4 ms after stimulation (not
shown). The barrel from which the activity originated was
never activated again through recurrent pathways, suggesting
that it was either still inhibited or in the refractory period when
the neighboring barrel was activated.
In two experiments, the signal remained in the home barrel
and did not propagate to other layers. In these cases, the
stimulation strength was probably too weak to reach threshold
for network activation and thus subsided before spreading to
L2/3.
Inhibition shapes the spatiotemporal evolution
of the columnar activity
As shown in Fig. 2, inhibition always follows and depends
on activation of the excitatory network. To study the effect of
the inhibitory network on the spatiotemporal evolution of the
excitation within the barrel cortex, we compared the activity
maps recorded under control condition and after the application
of 10 ␮M bicuculline (n ⫽ 11). Under control condition (Fig.
5A1), the spatial and temporal evolution of excitation and
91 • APRIL 2004 •
www.jn.org
Downloaded from http://jn.physiology.org/ by 10.220.33.5 on June 18, 2017
FIG. 3. Barrels are independent entities. A: bright
field microscope image of a thalamocortical slice
placed on the MEA where 3 barrels are clearly visible (black). The white square indicates a questionable barrel. B: activity maps recorded at the time
after stimulation indicated above each panel. Stimulus location is marked by a black dot. Early excitation is limited to regions corresponding primarily to
the barrels outlined in A (black). One barrel (white
square) that cannot be seen in A shows up in the
activity maps. Note that time of maximal activation
is different for each barrel. C: the location of the
centers of mass and their SDs were plotted as a
function of time after stimulation (see METHODS).
They remained horizontally confined to a small region of ⬃200 ␮m corresponding to the width of the
stimulated barrel (9 different experiments). D: superimposed activity profiles across barrels (thin black
traces, 10 experiments). The average of these 10
traces (thick black trace) was fitted with a Gaussian
curve (red) having a SD of 95 ␮m. The maximal
voltage was normalized to 1 in each experiment. E:
the centers of mass (average over 1.5–5 ms after
stimulation) of the positive signals and their SDs were
calculated from 5 different experiments. In each experiment L4 was stimulated at discrete locations along a
horizontal line, as shown in B. The distribution of the
center of mass along the MEA columns showed discrete clusters corresponding to single barrels. Each
color indicates a different experiment.
1640
C. WIRTH AND H.-R. LÜSCHER
inhibition develops as described in the previous section. In Fig.
5B1, the same experiment as shown in A1 is repeated under 10
␮M bicuculline. Now, inhibition is completely abolished, and
excitation is longer-lasting in L4 and lower L2/3 compared
with control conditions. Under bicuculline, the excitation wave
only rarely reached the uppermost L2/3.
The summary plots in Fig. 5, A2 and B2, show the spatial
and temporal evolution of the center of mass of the excitation
under control condition (A2) and under bicuculline (B2). Comparing these two figures reveals that the duration of excitation
in L4 and lower L2/3 was markedly increased after the application of bicuculline. With inhibition blocked, excitation remained 2–5 ms longer in the stimulated barrel and in lower
L2/3. In addition, the positive signal did not propagate to the
upper L2/3 as observed under control conditions. In control
conditions, the fast-moving inhibition starting from the barrel
and spreading to lower L2/3 “displaced” the positive signal
upward, limiting the duration of excitation. In upper L2/3,
excitation was substantially longer but restricted to the uppermost L2/3. To illustrate in more detail the influence of bicuculline on the network activity evoked by barrel stimulation,
we subtracted the recordings made under bicuculline from the
recordings taken under control conditions (Fig. 5C1). The so
calculated “inhibitory signals” in L4 started 3.2 ms after stimJ Neurophysiol • VOL
ulation and quickly moved up into lower L2/3. This strong
inhibitory, long-lasting signal was observed in all experiments
(n ⫽ 11). It never propagated in upper L2/3.
On the contrary, a sharp and laterally spreading small band
of a weak positive signal developed in upper L2/3 in parallel to
the negative signal located below. The temporal evolution of
the center of mass of the negative signals in L4 and L2/3 (Fig.
5C2, red dots, n ⫽ 7) shows a progression from L4 to lower
L2/3 in all experiments. This reliable, fast, and strong inhibition was responsible for shaping the diffuse excitation in L4
and L2/3 illustrated in Fig. 5B, 1 and 2, under bicuculline into
a sharp and narrow excitation window in upper L2/3 shown
under control conditions in Fig. 5A, 1 and 2.
The lateral spread of excitation seemed not to be increased
in L4 after inhibition was blocked. To evaluate the lateral
signal spread before and after the addition of bicuculline, we
calculated the activity profile across a barrel as in Fig. 3D in
control conditions and under bicuculline. The mean of the
activity profiles from 10 different experiments are plotted in
Fig. 6A and were fitted with a Gaussian curve yielding SDs of
70 ␮m (control) and 60 ␮m (bicuculline), respectively. This
difference was not statistically significant (P ⬎ 0.975, t-test on
the individual SDs). It suggests that inhibition is not responsible for the lateral confinement of the excitatory signals to a
91 • APRIL 2004 •
www.jn.org
Downloaded from http://jn.physiology.org/ by 10.220.33.5 on June 18, 2017
FIG. 4. Spatiotemporal evolution of activity
after stimulation of a single barrel. A1: time
series of activity maps spaced at an interval of
0.4 ms (red indicates depolarization; blue indicates hyperpolarization). Excitatory signals were
initiated in the stimulated barrel and then spread
to L2/3. Excitation mostly respected the columnar organization in L4 and lower L2/3. Excitation spread in horizontal direction in the upper
L2/3, resulting in a mushroom-like activation of
L2/3. Excitation was followed by inhibition
flowing from the stimulated barrel to the lower
L2/3. A2: mean activity over the MEA columns
covering the stimulated cortical column as a
function of time. This “line scan” of the activity
across the cortical layers of the same experiment
shown in A1 illustrates the spatiotemporal evolution of excitation moving from L4 to L2/3.
Excitation in L4 and lower L2/3 was quickly
replaced by a long-lasting inhibition. Black point
indicates location and time of stimulation, arrow
indicates the stimulation artifact. Excitation in
the upper-most layer remained active for ⬃10 ms
after stimulation. B: consecutive activity maps of
a different experiment after stimulation in a barrel (black dot) illustrating activation of a neighboring barrel through a L4-L2/3-L4 pathway. C:
same as in B but illustrating activation of a
neighboring barrel as well as of L5/6.
SPATIOTEMPORAL EVOLUTION OF ACTIVITY IN BARREL CORTEX
1641
Downloaded from http://jn.physiology.org/ by 10.220.33.5 on June 18, 2017
FIG. 5. The inhibition shapes the excitation in space and time: The temporal evolution of activity recorded under ACSF (A) and
under bicuculline (B) as well as the GABAA-mediated component (C). A,1 top: under ACSF, excitation was initiated in the barrel
and spread to lower L2/3. It was replaced by inhibition 3.4 ms after stimulation. Bottom: line scan of the same experiment
illustrating the long-lasting inhibition in L2/3 restricting excitation to the upper-most layer. Location and time of stimulation is
indicated by a black dot. A2: the centers of mass of the excitatory signals in L2–L4 were calculated (7 superimposed experiments)
and plotted as a function of time after stimulation. It illustrates the shift of excitation from L4 to L2/3 and the confinement of the
excitatory signal to a narrow band in L2/3 at later times. Horizontal line indicates the layer border. B1: same experiment as in A
after addition of 10 ␮M bicuculline. Excitation remains much longer in L4 and lower L2/3, whereas upper L2/3 is not excited. B2:
plot of center of mass of excitatory signals under bicuculline (6 experiments) shows the long-lasting and more diffuse activation
of L2/3. C1: subtracting the recordings in B1 from those in A1 yields the evolution of the signals due to GABAA-receptor activation,
revealing early, weak inhibition in the barrel and a later long-lasting inhibition in lower L2/3. In addition to the long-lasting
inhibition in L4 and lower L2/3 it illustrates the excitatory action on the upper-most layer. C2: plotting the center of mass of the
negative signals of the difference between recordings under bicuculline and recordings under ACSF (red trace, 7 experiments)
illustrates that inhibition stops excitation in L4 and lower L2/3 and constrains it to the upper L2/3. For comparison, the data from
A2 are included (blue dots). Different color codes were used in A2 and C2 to optimally display the changes in excitation and
inhibition, respectively.
J Neurophysiol • VOL
91 • APRIL 2004 •
www.jn.org
1642
C. WIRTH AND H.-R. LÜSCHER
barrel in L4. In addition we found that the lateral extent of
excitation in the lower part of L2/3 under bicuculline did not
increase either (Fig. 6B same 10 experiments as in Fig. 6A,
same SD of 105 ␮m). However, under both conditions (with
and without bicuculline) the response was always significantly
broader in L2/3 than in L4 (P ⬍ 0.001 with bicuculline, and
P ⬍ 0.025 without bicuculline).
Temporal evolution of the activity map after direct
stimulation of L6
In an additional set of experiments, we investigated L6 –L4
connections by stimulating L6, a second input layer, vertically
below a barrel above the white matter border (n ⫽ 15). Before
this was done, the barrel was always stimulated directly to
verify its activation pattern as shown in the previous sections.
Figure 7 illustrates a typical example of the spatiotemporal
evolution of the activity after L6 stimulation. In 13 of 15
experiments, stimulation in L6 resulted in a very early wave of
inhibition starting in L6 already 1 ms after stimulation. About
1.5 ms after stimulation, a first excitation wave was visible in
L4 usually confined to a barrel border. This excitation was
extremely short lasting (⬃0.5 ms) because it was suppressed
J Neurophysiol • VOL
DISCUSSION
Technical considerations
Multielectrode arrays (MEAs) are a new tool for recording
the two-dimensional spread of activity in acute brain slices
with high temporal resolution.
The design of the MEA and thus the spatial resolution is
limited by the number of available amplifiers for simultaneous
91 • APRIL 2004 •
www.jn.org
Downloaded from http://jn.physiology.org/ by 10.220.33.5 on June 18, 2017
FIG. 6. Bicuculline does not influence the lateral spread of excitation in L4
and lower L2/3. A: normalized and averaged (n ⫽ 10 experiments) excitation
profile across barrels in L4 under control condition (red circles) and under 10
␮M bicuculline (blue circles). The traces are fitted with Gaussian curves
yielding SDs of 70 ␮m (control, red trace) and 60 ␮m (bicuculline, blue trace),
respectively. There is no significant difference between the 2 curves. B: the
same analysis as shown in A, but the excitation profiles were calculated across
the cortical columns in L2/3 (same 10 experiements as in A). Again no
significant difference in the width of the excitation profiles were seen under
control conditions and under bicuculline (SD 105 ␮m under both conditions).
However, the excitation profile was significantly broader in L2/3 than in L4.
by the rising inhibition from L5/6. This inhibition wave rose
from L5/6 to L4 and further to lower L2/3 before it was
replaced in L4 by an excitation wave (starting 2.2 ⫾ 0.9 ms in
L6, 14 experiments), which propagated from L6 to L4 within
⬃2 ms, activating sometimes a single barrel and sometimes L4
in a more diffuse manner. Similarly as after direct stimulation
in L4, activation in L4 was short lasting (1.9 ⫾ 1.1 ms, n ⫽
14). From L4, excitation spread to lower L2/3 as described in
the previous section. Simultaneously, excitation flowed back
from L4 to L6, sometimes through a neighboring barrel, fanning out into a broad band of excitation in L6, including the
cortical column from which the excitation wave started. Excitation in L4 was followed by strong inhibition that remained
active for several ms (5.5 ⫾ 2.7 ms, n ⫽ 14) but hardly spread
to supragranular layers.
The time course of the center of mass of the positive response (Fig. 8A, 7 experiments) revealed a clear L6-L4-L6
loop in all experiments. In one experiment, the excitation did
not reach L4 but took a short cut through L5, thus following a
L6-L5-L6 loop (not shown). Similarly to direct barrel stimulation, the spatiotemporal evolution of the inhibitory signal was
more stereotypical compared with the excitatory signals after
stimulation of L6. Figure 8B illustrates the temporal evolution
of the center of mass of the negative wave. Inhibition ascended
like excitation to L4, where it stayed active for ⱕ10 ms.
(Although inhibition was shortly “interrupted” by the ascending excitation between 3.5 and 4.5 ms in individual experiments, this “brake” in the inhibition is not visible in Fig. 8B
because several experiments are superimposed and excitation
occurred at slightly different times). Inhibition never followed
the excitatory loop back into L6.
In Fig. 8C, the temporal evolution of the center of mass of
the positive signal after stimulation in L6 is plotted versus the
lateral distance (7 experiments superimposed) and confirms the
preceding observation that activation of L4 through the L6 to
L4 pathway did not always respect the barrel borders. Even
when excitation was confined within the barrel borders for the
first 3 ms after stimulation in L6, it crossed the barrel borders
for later activation time. In contrast the inhibitory wave always
respected the barrel border even 10 ms after stimulation in L6
(Fig. 8D).
To elucidate the interplay between excitatory and inhibitory
signals after L6 stimulation, we compared the activity maps
recorded under control condition and after the application of 50
␮M D-AP5 and 10 ␮M CNQX (n ⫽ 3, not shown). Addition of
D-AP5 and CNQX abolished all excitation and late inhibition.
However the early ascending negative signals remained mostly
unchanged. After adding 10 ␮M bicuculline, this early negative signal disappeared as well. Thus the early ascending
inhibition must be caused by direct stimulation of inhibitory
fibers that ascend to L4.
SPATIOTEMPORAL EVOLUTION OF ACTIVITY IN BARREL CORTEX
1643
recording, the tip height of the electrodes, and the surface area of
the tissue to be covered by the MEA. The spacing between the
electrodes could be reduced with a different MEA-design by at
least a factor of 2; however, this would decrease the tip height of
the electrodes as well as the total covered surface area (using the
same number of electrodes) (Heuschkel et al. 2002). Thus we used
a rectangular array with interleaved electrodes between adjacent
rows, an interspacing of 140 ␮m and a tip high of 42 ␮m, that
allowed us to record over the entire thickness of the somatosensory cortex with a relatively high spatial resolution and still
covering three to four barrels across the barrel field.
Compared with voltage-sensitive dye imaging with highresolution CCD imaging devices (Laaris and Keller 2002;
Petersen and Sakmann 2001), the spatial resolution of the
MEA is therefore more limited. On the other hand, MEA data
can be sampled at a high rate of 10 kHz, outperforming the
limited sampling rate of the CCD cameras. The quality of LFP
recordings with MEAs depends on secure contact between the
individual electrodes and the brain tissue, which we have
optimized by coating the electrodes with PEI and by using
pyramidal shaped instead of planar electrodes.
Interpretation of the LFP recordings
In principle, LFPs reflect the voltage drop produced by
current flow across the extracellular resistance. The current
J Neurophysiol • VOL
sources or sinks have their origin in local synaptic conductances and propagated action potentials. We did not calculate
the current source density profiles because of the low spatial
resolution and uncertainties in the direction of current flow (for
review, see Mitzdorf 1985). We reduced instead the complexity of the LFP-recordings by subtracting the signal produced by
fiber volleys and terminal potentials from the LFP-recordings
(see METHODS). The underlying source of the reduced LFP was
limited to synaptic currents.
By means of pharmacological interventions we could clearly
identify that the late negative sLFP was due to activation of
GABAA receptors. The early positive component of the sLFP
was mainly due to activation of AMPA receptors. In addition
a substantial contribution of NMDA receptor activation could
be seen in agreement with findings by Egger et al. (1999).
Application of CNQX and D-AP5 not only blocked the early
positive component of the LFP but consistently also abolished
the late negative component. This is consistent with the observation of thalamocortical spike-triggered rabbit LFPs, which
completely disappeared after the local application of DNQX
(Swadlow and Gusev 2000). We therefore conclude that the
positive component of the sLFP is the result of stimulating
mainly excitatory TC fibers and excitatory axons from L4 cells,
which evoke glutamatergic postsynaptic potentials on excitatory as well as on inhibitory neurons in L4 barrels. The nega-
91 • APRIL 2004 •
www.jn.org
Downloaded from http://jn.physiology.org/ by 10.220.33.5 on June 18, 2017
FIG. 7. Response pattern after stimulation of L6. A1: time series of activity maps showed a very early inhibition ascending from
L5/6 to a barrel in L4. It is followed by raising excitation flowing from L6 to L4, subsequently to L2/3 and simultaneously back
to L6. Excitation in L4 is followed by inhibition that may be caused by interneurons activated through excitatory L4 neurons. Black
dot: Stimulation position. A2: line scan of the same experiment shown in A1, illustrating the ascending inhibition followed by an
excitatory L6-L4-L6 loop. Excitation flows to L2/3 as well. In L4 excitation is replaced by a long-lasting inhibition at ⬍4 ms after
stimulation.
1644
C. WIRTH AND H.-R. LÜSCHER
tive component results from feedforward as well as from
feedback inhibition through GABAergic synapses in agreement with previous findings (Bruno and Simons 2002; Gibson
et al. 1999; Porter et al. 2001; Swadlow 2003). Occasionally,
the stimulus could activate directly a small population of
inhibitory neurons as evidenced by a small early persistent
negativity that could be blocked by the application of bicuculline (see Fig. 2C). Although we could clearly relate positivity
and negativity in the sLFP to the corresponding local synaptic
excitatory or inhibitory actions that underlay active current
sources or sinks, we cannot exclude the possibility that the
sLFP is also influenced by passive currents sources and sinks
produced by longitudinal current flow within the long apical
dendrites of pyramidal neurons. For example, the long-lasting
narrow band of positivity in upper L2/3 (Fig. 5A1, bottom)
could reflect the passive current sink produced by the strong
inhibition in lower L2/3. On the other hand, the positive signal
in upper L2/3 started before the strong inhibition developed
below in lower L2/3, and inhibition seemed to end before the
“excitation” in upper L2/3. A similar line of argumentation
could be followed for the weak negativity seen in upper L2/3
under bicuculline (Fig. 5B1). Here again, the negativity developed only after the positivity starts to decline again.
These difficulties are not restricted to the interpretation of
LFPs but would remain the same for the interpretation of
current source-density profiles.
with bicuculline, the early excitatory signals in the barrels
respected the barrel borders, demonstrating that excitation is
not spatially constrained to a single barrel by inhibitory networks. This finding confirms previous studies with voltagesensitive dyes (Laaris and Keller 2002; Petersen and Sakmann
2001) and is supported by recent anatomical and physiological
studies showing that axonal and dendritic arbors in L4 are
largely confined to individual barrels (Feldmeyer et al.1999;
Lübke et al. 2000; Petersen and Sakmann 2000; Staiger et al.
1999). Direct connections between neurons of different barrels
are very rare (Kim and Ebner 1999; Schubert et al. 2003).
Excitation of neighboring barrels always occurred via L2/3 and
never through direct communication between barrels. A weak
direct excitation reported by Schubert et al. (2003) might not
be detectable in field potential recordings.
Subsequent spread of excitation to L2/3 in a mushroom-type
manner has also been described previously (Feldmeyer et al.
2002; Petersen and Sakmann 2001). Petersen and Sakmann
(2001) observed enlarged lateral spread of excitation in L2/3 of
thalamocortical slices after blocking inhibition with 10 ␮M
bicuculline. In contrast, we did not see a significant expanded
lateral spread of excitation in supragranular layers after blocking inhibition. The limited spatial resolution of MEAs might
prevent detection of a small increase in signal spread after
application of bicuculline.
Role of the inhibitory network
Functional neocortical column
The direct observation of the spatiotemporal spread of excitation and inhibition through all laminae of the cortex provides a dynamic view of the interdependence of excitation and
inhibition in the cortical column. Even after blocking inhibition
J Neurophysiol • VOL
The results presented reveal a strong and rather stereotypic
spatial and temporal interplay between excitatory and inhibitory signals in the barrel cortex. Inhibition occurs mainly
through the excitatory network and not by direct electrical
stimulation, reminiscent of feedback inhibition. This observa-
91 • APRIL 2004 •
www.jn.org
Downloaded from http://jn.physiology.org/ by 10.220.33.5 on June 18, 2017
FIG. 8. Evolution of center of mass of excitation and inhibition after stimulation in L6. An excitatory (A and C) and
an inhibitory (B and D) pathway with different spatiotemporal
behavior are activated. A: the evolution of the center of mass
of the excitatory signals in L4-L6 reveals an excitatory L6L4-L6 loop, as observed in Fig. 7A. Because the time course
of the L6-L4-L6 was different in different experiments, the
total loop time was normalized before the data points of the
different experiments (n ⫽ 7) were superimposed. B: the
center of gravity of the inhibitory signal in L5/6 to L4 is
plotted against time after stimulation. Inhibition ascends
within ⬃3 ms in a stereotypical manner from L6 to L4. The
short excitatory brake in the inhibitory signal (compare Fig.
7A2) cannot be seen in this illustration, because this brake
occurs at different times in the individual experiments and is
therefore blurred in this illustration. C: the centers of mass of
excitatory signals in L4 are plotted against time after stimulation, demonstrating that the barrel borders were not respected if excitation reached the barrel through the L6 pathway. D: the centers of mass of inhibitory signals in L4 are
plotted against time after stimulation. The inhibitory signals
in L4 remain confined to barrel borders after it rose from L6
to L4. The results of 7 experiments are shown in each panel.
SPATIOTEMPORAL EVOLUTION OF ACTIVITY IN BARREL CORTEX
Response in L6
Contrary to the response after stimulation of single barrels,
excitatory and inhibitory networks seem to be activated independently after stimuli in L6, indicating two separate circuits
with distinct functions. This observation is in agreement with
findings in monkey visual cortex where multiple, distinct local
circuits are suggested in infragranular layers (Briggs and Callaway 2001).
The early stereotype negative signal spreading to L4 and
preceding excitation is surprising. It occurred only 1 ms after
stimulation and remained after CNQX and D-AP5 application.
This suggests that inhibitory fibers were stimulated directly and
are probably engaged in feedforward inhibition in L6 as well as
to L4. Inhibition of L4 cells originating in L6 has been obJ Neurophysiol • VOL
served in the visual cortex (Bolz and Gilbert 1986; Hirsch
1995) and has been implicated in the generation of specific
receptive field properties such as end-inhibition. We propose
that a similar inhibitory L6-L4 pathway is present in rat somatosensory cortex probably controlling duration and strength
of excitation in the barrels.
Strong excitatory signals could only develop when inhibition
in L6 faded away. This slow and diffuse excitation suggests a
more complicated circuit with more synapses involved than the
ascending inhibitory pathway. It is possible that TC fibers
passing through L6 were directly stimulated. Two different TC
cells have been identified (Arnold et al. 2001): Direct-projecting axons that arborize within a single barrel and bifurcatingtype axons that bifurcate in the white matter or in L6 and then
project to multiple barrels. This could explain the heterogeneous activation pattern observed after L6 stimulation leading
to both, single-barrel excitation as well as to a more diffuse
activation of several barrels. It is also possible that L6-L4
fibers were directly stimulated. Indeed, excitatory L6-L4 connections have been found in rat somatosensory cortex (Laaris
and Keller 2002; Staiger et al. 1996; Zhang and Deschênes
1997) as well as in cat visual cortex (Bolz and Gilbert 1986;
Stratford et al. 1996; Tarczy-Hornoch et al. 1999). In addition,
direct input from barrel neurons to infragranular layers below
the parent barrel has been observed in rat barrel cortex (Armstrong-James 1995; Laaris and Keller 2002; Schubert et al.
2001) and cat visual cortex (Tarczy-Hornoch et al. 1999). An
excitatory L6-L4-L6 loop in rat barrel cortex is thus probable
and has been proposed by Lübke et al. (2000). Our findings are
an additional indication that such an excitatory feedforward
loop exists. It may contribute to a strong excitation in L4 and
may provide L6 with a feedback information on the state of
excitation in L4.
Functional implications and conclusions
In the freely exploring rat, whisking occurs at a frequency of
⬃8 –10 Hz. Well-trained rats are able to discriminate between
a smooth surface and one having shallow grooves spaced at
intervals of 50 ␮m. During forward whisker sweeps, the hair
moves across the surface at a speed of 10 –20 mm/s (for a
review, see Simons 1995). Therefore the individual whisker
senses surface perturbation at a frequency of once every 2.5–5
ms. The very short window of excitation seen in L4 and lower
L2/3 might be necessary to accommodate this high frequency
in the barrel cortex. The excitation-inhibition cycle seen was
however longer than the 5 ms suggested by the preceding
simple calculations. It is well possible that during repetitive
activation, the rising excitation might be able to override the
inhibition due to differences in short-term plasticity (Fuhrmann
et al. 2002). The close interaction between excitation and
inhibition would enable the barrel network to follow a fast
cycle of excitation and inhibition. Interestingly this short window of excitation is restricted to L4 and lower L2/3. In contrast
to this short excitation, a longer-lasting and sharply delineated
excitatory band is established in the upper L2/3. One could
speculate that this excitatory action impinging on the distal
dendrites might prepare L5 pyramidal cells, the output neurons
of the cortical column, for top-down interactions by lowering
the threshold for backpropagation activated Ca2⫹ spike firing
(BAC firing) (Larkum et al. 1999). The two windows of
91 • APRIL 2004 •
www.jn.org
Downloaded from http://jn.physiology.org/ by 10.220.33.5 on June 18, 2017
tion probably reflects the fact that only ⬃15% of all neuronal
somata in the barrel belong to inhibitory neurons and are thus
not likely to be stimulated directly (Beaulieu 1993).
Flow of inhibition from L4 to supragranular layers after
activation of L4, as reported here, has been predicted by Porter
et al. (2001). They observed that axonal arbors of L4 interneurons are mostly directed upwardly, and he suggests that inhibition follows excitation along its intracortical course, being
inseparably linked to it. This prediction is confirmed by our
observation, demonstrating that the rising inhibition from L4 to
L2/3 always follows excitation and, more importantly, curtails
excitation to a few milliseconds. The results reported in this
paper reveal a fundamental feature of the inhibitory network
with regard to shaping the excitation not only in the temporal
but also in the spatial domain. The temporal aspect of shaping
the excitatory network has been described for L4 where synchronous discharge of interneurons after thalamocortical excitatory input generated a synchronous feedforward inhibition
within the barrel, allowing only a brief window of excitability
(Swadlow 2003). Miller et al. (2001) proposed a strong recurrent excitatory barrel circuit reinforcing the thalamocortical
excitation rapidly and nonlinearly, which is then stopped after
a few milliseconds by feedforward inhibition that grows more
linearly and thus develops too late to prevent the initiation of
an excitatory response but curtails excitation efficiently. This is
in agreement with our findings that excitation in a barrel is cut
off after a few milliseconds by a rise in inhibition. However,
ascending inhibition stops excitation in L2/3 as well. This is
demonstrated by the long-lasting activation of lower L2/3 with
blocked GABAA receptors. In addition to the temporal shaping
of the excitatory response in the barrel cortex, a prominent
spatial shaping of the response could be demonstrated as well.
With inhibition present, the more prominent positive signal in
the uppermost L2/3 most likely reflects an excitatory action of
GABA. As shown by Gulledge and Stuart (2003), GABA
responses in the distal apical dendrite were always excitatory.
Our results confirm their findings by showing that inhibition
alone produces a negative signal in L4 and lower L2/3 but a
strong positive signal in upper L2/3.
The excitatory signals often varied in space and time, but the
response of the inhibitory network (its excitatory action as
well) was very stereotypical in all experiments, emphasizing
the enormous importance of a precise and solid inhibition
which shapes the varying excitatory response in time and space
to enable a reliant and precise excitation in the barrel cortex.
1645
1646
C. WIRTH AND H.-R. LÜSCHER
excitation created in the apical dendritic tuft and the basal
dendrite (see Fig. 9) would put the L5 pyramidal neuron into
the unique position to associate top-down input through BAC
firing (Larkum et al. 1999) from motor cortical areas and
bottom-up input from the whisker pad leading to behaviorally
relevant output from the somatosensory cortex.
ACKNOWLEDGMENTS
We thank Drs. Berger, Senn, Larkum, and Petersen for helpful discussions
and careful reading of an earlier version of this manuscript.
GRANTS
This work was supported by the Swiss National Science Foundation, Grant
3100-061335.00, and the Silva Casa Foundation to H.-R. Lüscher.
REFERENCES
Agmon A and Connors BW. Thalamocortical responses of mouse somatosensory (barrel) cortex in vitro. Neuroscience 41: 365–379, 1991.
Amitai Y, Gibson JR, Beierlein M, Patrick SL, Ho AM, Connors BW, and
Golomb D. The spatial dimensions of electrically coupled networks of
interneurons in the neocortex. J Neurosci 22: 4142– 4152, 2002.
Armstrong-James M, Fox K, and Das-Gupta A. Flow of excitation within
rat barrel cortex on striking a single vibrissa. J Neurophysiol 68: 1345–1358,
1992.
Armstrong-James M. The nature and plasticity of sensory processing within
adult rat barrel cortex. In: The Barrel Cortex of Rodents, edited by Jones EG
and Diamond IT. New York: Plenum, 1995, p. 333–373.
Arnold PB, Li CX, and Waters RS. Thalamocortical arbors extend beyond
single cortical barrels: an in vivo intracellular tracing study in rat. Exp Brain
Res 136: 152–168, 2001.
Bacci A, Huguenard JR, and Prince DA. Functional autaptic neurotransmission in fast-spiking interneurons: a novel form of feedback inhibition in the
neocortex. J Neurosci 23: 859 – 866, 2003.
Beaulieu C. Numerical data on neocortical neurons in adult rat, with special
reference to the GABA population. Brain Res 609: 284 –292, 1993.
Beierlein M and Connors BW. Short-term dynamics of thalamocortical and
intracortical synapses onto layer 6 neurons in neocortex. J Neurophysiol 88:
1924 –1932, 2002.
Beierlein M, Gibson JR, and Connors BW. A network of electrically
coupled interneurons drives synchronized inhibition in neocortex. Nat Neurosci 3: 904 –910, 2000.
Bolz J and Gilbert CD. Generation of end-inhibition in the visual cortex via
interlaminar connections. Nature 320: 362–365, 1986.
J Neurophysiol • VOL
91 • APRIL 2004 •
www.jn.org
Downloaded from http://jn.physiology.org/ by 10.220.33.5 on June 18, 2017
FIG. 9. Summary diagram of the temporal and spatial evolution of activity
through the cortical layers. The spatial spread of the activity along the horizontal axis cannot be shown in this diagram. The time evolution occurs along
the abscissa. The arrival time of the TC input in L6 is defined as time 0. Short
horizontal arrows indicate synaptic delays. Excitatory pathways are indicated
with solid lines, inhibitory pathways with dotted lines. Light gray shading
indicates excited regions, whereas dark gray shading indicates inhibited regions. The vertical short and dotted arrows in L2/3 indicate excitatory action
of the GABAergic inhibition. One L5 pyramidal cell is schematically indicated. Inhibition opens two windows of excitation, one in L5/6 and one in the
upper L2/3. The L5 output neuron may “feel” these two windows of excitation
preparing the cell for associating top-down input from other cortical areas and
bottom-up input from peripheral sensory pathways. Further explanation in the
text.
Bruno RM and Simons DJ. Feedforward mechanisms of excitatory and
inhibitory cortical receptive fields. J Neurosci 22: 10966 –10975, 2002.
Briggs F and Callaway EM. Layer-specific input to distinct cell types in layer
6 of monkey primary visual cortex. J Neurosci 21: 3600 –3608, 2001.
Diamond ME. Somatosensory thalamus of the rat. In: The Barrel Cortex of
Rodents, edited by Jones EG and Diamond IT. New York: Plenum, 1995, p.
189 –219.
Egert U, Heck D, and Aertsen A. Two-dimensional monitoring of spiking
networks in acute brain slices. Exp Brain Res 142: 268 –274, 2002.
Egger V, Feldmeyer D, and Sakmann B. Coincidence detection and changes
of synaptic efficacy in spiny stellate neurons in rat barrel cortex. Nat
Neurosci 2: 1098 –1105, 1999.
Feldmeyer D, Egger V, Lübke J, and Sakmann B. Reliable synaptic connections between pairs of excitatory layer 4 neurons within a single “barrel”
of developing rat somatosensory cortex. J Physiol 521: 169 –190, 1999.
Feldmeyer D, Lübke J, Silver RA, and Sakmann B. Synaptic connections
between layer 4 spiny neuron - layer 2/3 pyramidal cell pairs in juvenile rat
barrel cortex: physiology and anatomy of interlaminar signaling within a
cortical column. J Physiol 538: 803– 822, 2002.
Fuhrmann G, Segev I, Markram H, and Tsodyks M. Coding of temporal
information by activity-dependent synapses. J Neurophysiol 87: 140 –148,
2002.
Galarreta M and Hestrin S. Electrical and chemical synapses among parvalbumin fast-spiking GABAergic interneurons in adult mouse neocortex.
Proc Natl Acad Sci USA 99: 12438 –12443, 2002.
Gibson JR, Beierlein M, and Connors BW. Two networks of electrically
coupled inhibitory neurons in neocortex. Nature 402: 75–79, 1999.
Goldreich D, Kyriazi HT, and Simons DJ. Functional independence of layer
IV barrels in rodent somatosensory cortex. J Neurophysiol 82: 1311–1316,
1999.
Gottlieb JP and Keller A. Intrinsic circuitry and physiological properties of
pyramidal neurons in rat barrel cortex. Exp Brain Res 115: 47– 60, 1997.
Gulledge AT and Stuart GJ. Excitatory action of GABA in the Cortex.
Neuron 37: 299 –309, 2003.
Heuschkel MO, Fejtl M, Raggenbass M, Bertrand D, and Renaud P. A
three-dimensional multielectrode array for multi-site stimulation and recording in acute brain slices. J Neurosci Methods 114: 135–148, 2002.
Hirsch JA. Synaptic integration in layer 4 of the ferret striate cortex. J Physiol
483: 183–199, 1995.
Jensen KF and Killackey HP. Terminal arbors of axons projecting to the
somatosensory cortex of the adult rat. I. The normal morphology of specific
thalamocortical afferents. J Neurosci 7: 3529 –3543, 1987.
Keller A. Synaptic organization of the barrel cortex. In: The Barrel Cortex of
Rodents, edited by Jones EG and Diamond IT. New York: Plenum, 1995, p.
221–262.
Kim U and Ebner FF. Barrels and septa: separate circuits in rat barrel cortex.
J Comp Neurol 408: 489 –505, 1999.
Kleinfeld D and Delaney KR. Distribute representation of vibrissa movement
in the upper layers of somatosensory cortex revealed with voltage-sensitive
dyes. J Comp Neurol 375: 89 –108, 1996.
Laaris N, Carlson GC, and Keller A. Thalamic-evoked synaptic interactions
in barrel cortex revealed by optical imaging. J Neurosci 20: 1529 –1537,
2000.
Laaris N and Keller A. Functional independence of layer IV barrels J Neurophysiol 87: 1028 –1034, 2002.
Larkum ME, Zhu JJ, and Sakmann B. A new cellular mechanism for
coupling inputs arriving at different cortical layers. Nature 398: 338 –341,
1999.
Lübke J, Egger V, Sakmann B, and Feldmeyer D. Columnar organization of
dendrites and axons of single and synaptically coupled excitatory spiny
neurons in layer 4 of the rat barrel cortex. J Neurosci 20: 5300 –5311, 2000.
Miller KD, Pinto DJ, and Simons DJ. Processing in layer 4 of the neocortical
circuit: new insights from visual and somatosensory cortex. Curr Opin
Neurobiol 11: 488 – 497, 2001.
Mitzdorf U. Current source-density method and application in cat cerebral
cortex: investigation of evoked potentials and EEG phenomena. Physiol Rev
65: 37–100, 1985.
Petersen RS and Diamond ME. Spatial-temporal distribution of whiskerevoked activity in rat somatosensory cortex and the coding of stimulus
location. J Neurosci 20: 6135– 6143, 2000.
Petersen CCH and Sakmann B. The excitatory neuronal network of rat layer
4 barrel cortex. J Neurosci 20: 7579 –7586, 2000.
SPATIOTEMPORAL EVOLUTION OF ACTIVITY IN BARREL CORTEX
J Neurophysiol • VOL
Staiger JF, Zilles K, and Freund TF. Connectivity in the somatosensory
cortex of the adolescent rat: an in vitro biocytin study. Anat Embryol (Berl)
199: 357–365, 1999.
Stratford KJ, Tarczy-Hornoch K, Martin KA, Bannister NJ, and Jack JJ.
Excitatory synaptic inputs to spiny stellate cells in cat visual cortex. Nature
382: 258 –261, 1996.
Swadlow HA. Fast-spike interneurons and feedforward inhibition in awake
sensory neocortex. Cereb Cortex 13: 25–32, 2003.
Swadlow HA and Gusev AG. The influence of single VB thalamocortical
inpulses on barrel columns of rabbit somatosensory cortex. J Neurophysiol
83: 2802–2813, 2000.
Tarczy-Hornoch K, Martin KA, Stratford KJ, and Jack JJ. Intracortical
excitation of spiny neurons in layer 4 of cat striate cortex in vitro. Cereb
Cortex 9: 833– 843, 1999.
Woolsey TA and Van der Loos H. The structural organization of layer IV in
the somatosensory region (SI) of the mouse cerebral cortex. The description
of a cortical field composed of discrete cytoarchitectonic units. Brain Res
17: 205–242, 1970.
Zhang ZW and Deschênes M. Intracortical axonal projections of lamina VI
cells of the primary somatosenory cortex in the rat: a single cell labeling
study. J Neurosci 17: 6365– 6379, 1997.
91 • APRIL 2004 •
www.jn.org
Downloaded from http://jn.physiology.org/ by 10.220.33.5 on June 18, 2017
Petersen CCH and Sakmann B. Functionally independent columns of rat
somatosensory barrel cortex revealed with voltage-sensitive dye imaging.
J Neurosci 21: 8435– 8446, 2001.
Pinto DJ, Hartings JA, Brumberg JC, and Simons DJ. Cortical damping:
analysis of thalmocortical response transformations in rodent barrel cortex.
Cereb Cortex 13: 33– 44, 2003.
Porter JT, Johnson CK, and Agmon A. Diverse types of interneurons
generate thalamus-evoked feedforward inhibition in the mouse barrel cortex,
J Neurosci 21: 2699 –2710, 2001.
Schubert D, Kötter R, Zilles K, Luhmann HJ, and Staiger JF. Cell type-specific
circuits of cortical layer IV spiny neurons. J Neurosci 23: 2961–2970, 2003.
Schubert D, Staiger JF, Cho N, Kötter R, Zilles K, and Luhmann HJ.
Layer-specific intracolumnar and transcolumnar functional connectivity of layer
V pyramidal cells in rat barrel cortex. J Neurosci 21: 3580–3592, 2001.
Simons DJ and Woolsey TA. Morphology of Golgi-Cox-impregnated barrel
neurons in rat SmI cortex. J Comp Neurol 230: 119 –132, 1984.
Simons DJ. Neuronal integration in the somatosensory whisker/barrel cortex.
In: The Barrel Cortex of Rodentse edited by Jones EG and Diamond IT. New
York: Plenum, 1995, p. 263–297.
Staiger JF, Zilles K, and Freund TF. Distribution of GABAergic elements
postsynaptic to ventroposteromedial thalamic projections in layer IV of rat
barrel cortex. Eur J Neurosci 8: 2273–2285, 1996.
1647