• No results found

Phase-dependent effects of stimuli locked to oscillatory activity in cultured cortical networks

N/A
N/A
Protected

Academic year: 2021

Share "Phase-dependent effects of stimuli locked to oscillatory activity in cultured cortical networks"

Copied!
7
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Phase-Dependent Effects of Stimuli Locked to Oscillatory Activity

in Cultured Cortical Networks

Jan Stegenga,

*

Joost le Feber, Enrico Marani, and Wim L. C. Rutten

Biomedical Signals and Systems Group, Institute for Biomedical Technology and Technical Medicine, University of Twente, Enschede, The Netherlands

ABSTRACT

The archetypal activity pattern in cultures of dissociated neurons is spontaneous network-wide bursting. Bursts

may interfere with controlled activation of synaptic plasticity, but can be suppressed by the application of stimuli at a sufficient

rate. We sinusoidally modulated (4 Hz) the pulse rate of random background stimulation (RBS) and found that cultures were

more active, burst less frequently, and expressed oscillatory activity. Next, we studied the effect of phase-locked tetani (four

pulses, 200 s

1

) on network activity. Tetani were applied to one electrode at the peak or trough of mRBS stimulation. We found

that when tetani were applied at the peak of modulated RBS (mRBS), a significant potentiation of poststimulus histograms

(PSTHs) occurred. Conversely, tetani applied at the trough resulted in a small but insignificant depression of PSTHs. In addition

to PSTHs, electrode-specific firing rate profiles within spontaneous bursts before and after mRBS were analyzed. Here,

signif-icant changes in firing rate profiles were found only for stimulation at the peak of mRBS. Our study shows that rhythmic activity

in culture is possible, and that the network responds differentially to strong stimuli depending on the phase at which they are

delivered. This suggests that plasticity mechanisms may be differentially accessible in an oscillatory state.

INTRODUCTION

Multi-electrode arrays (MEAs) offer a means to investigate

synaptic plasticity on the small-network scale (

1–3

). Cultures

of dissociated cortical neurons create a monolayer of cells

on the MEA surface that are easily accessible for recoding

and stimulation, and may facilitate learning and memory

studies (

4

). The occurrence of spontaneous, synchronized

bursting in these nonstructured networks has made it difficult

to achieve consistent results on plasticity using dissociated

cultures (

5–10

). Bursts of action potentials, characteristic

for networks of dissociated neurons, resemble the type of

activity that is observed during early development of the

nervous system. This activity subsides as the brain starts to

receive input from sensory neurons (

11–14

); thus, a lack of

afferent external input to cultures may cause bursting. The

fact that bursts can be suppressed by random background

stimulation (RBS) supports this view (

7

). Given the high rate

of action potential firings, and the fact that NMDA receptors

are activated during bursts, it is reasonable to assume that

plasticity mechanisms are activated in an uncontrolled

man-ner (

15,16

). Furthermore, it has been shown that plasticity

mechanisms are more accessible by stimuli when bursts

are suppressed (

17–19

). Thus, the ability to suppress bursts

in cortical cultures may be beneficial for accessing and

assessing synaptic plasticity.

In this study, we took inspiration from the rhythmic

activity observed in the hippocampus, which is known to

be involved in long-term memorization (

20–24

) and learning

tasks (

22,25

). In particular, applying a train of stimuli (four

pulses, 200 s

1

) at the peak of hippocampal oscillation

results in long-term potentiation (LTP), whereas the same

train of stimuli applied at the trough results in long-term

depression (LTD) of postsynaptic potentials (

22,23

). One

hypothesis about the mechanism involved is that the

oscilla-tion of inhibitory neurons modulates the excitability of

neu-rons that are postsynaptic to the stimulated neuron and

consequently modulate the direction of change (

25

).

Oscilla-tory activity, regardless of its origin, may thus have a

pro-found effect on the way in which stimuli are processed by

the network.

In this study, we modified the RBS algorithm by using

stimuli that were Poisson-distributed in time (average of 10

stimuli per second) and delivered to a fully randomized

sequence of locations. Inspired by the oscillatory

depen-dency of plasticity in vivo, we sinusoidally (4 Hz) modulated

the Poisson parameter. By doing so, we found that the

appli-cation of rhythmically modulated RBS (mRBS) in culture

serves two purposes. First, it suppresses bursts that would

otherwise obscure induced plasticity, and evokes oscillating

activity in the network. Second, it modulates excitability in

the culture such that phase-locked trains of stimuli have

effects that are phase-specific.

We assessed changes in the network using two different

methods. First, we considered the response to probe stimuli.

We found that changes in the magnitude of responses were

pathway-specific and indeed depended on the phase at which

the tetani were delivered. Second, we analyzed spontaneous

bursting activity. Previous studies have shown that the

spatiotemporal patterns of activity during bursts are stable

over periods of several hours (

26–28

) but can also be

changed by proper electrical stimuli (

29–31

).

Submitted July 28, 2009, and accepted for publication February 16, 2010. *Correspondence:j.stegenga@utwente.nl

Editor: Herbert Levine.

(2)

MATERIALS AND METHODS

Culturing

Cells were obtained from the neocortices of newborn Wistar rats. The cells were dissociated mechanically by trituration, and chemically by treatment with trypsin. MEAs produced by Multi Channel Systems (Reutlingen, Germany) were coated with poly-ethyl-imine to promote cell-to-surface adhe-sion. Next, a drop of plating fluid was placed in the center of the MEA and the cells were allowed to attach for 4 h, after which the plating medium was washed and nonadhering cells were removed. The plating concentration was 1 million cells per milliliter, which resulted in a monolayer of cells with a density of ~2500 cells per mm2at 2 days in vitro (DIV). We used Romijn’s chemically defined R12 medium (32) for plating, maintenance, and experimentation. Half of the medium was replaced every 2 days. Cultures were stored in an incubator at 37C, 5% CO2, and near 100% humidity. The

platinum-nitride MEA electrodes were 30 mm in diameter and spaced 200 mm apart in an 88 grid (excluding the corner electrodes). The electrodes were numbered in matrix form (e.g., electrode 45 was located at the fourth column, fifth row). A total of eight different cultures, taken from five different platings, were used. Culture ages ranged from 20 to 89 DIV, with most (13/17) recording sessions performed between 20 and 35 DIV.

Experimental setup

Measurements

We built our measurement setup around a commercially available MEA-recording setup (1060BC preamplifier, STG1002 stimulus generator) from Multi Channel Systems. Data were acquired at a 16 kHz sampling rate with the use of a 6024E card (National Instruments, Austin, TX), and controlled by custom LabView programs (National Instruments). During measurements, the MEAs were sealed with a semipermeable membrane (Multi Channel Systems) and the temperature was controlled at 36C. The CO

2level was

maintained at 5% and the culture chamber was heated from the top by a Peltier element to prevent condensation of the medium. Cultures were left to acclima-tize in the experimental setup for 30 min before experiments were initiated. The stimuli were always monopolar, biphasic (200 ms per phase, positive phase first) current pulses. During mRBS, the same pulse was used for all electrodes. During probing sessions and spontaneous recordings, spikes were detected using a threshold crossing algorithm and validated in real time using an algo-rithm described by Wagenaar et al. (33). During mRBS sessions, continuous traces of all 60 electrodes were saved because spike validation interfered with the timing accuracy of stimuli. These data files were analyzed offline in the same way as would otherwise have been done in real time.

Protocol

All 60 electrodes were randomly probed four times with three different amplitudes (between 8 and 20 mA) at 5 s intervals to determine the subset of electrodes to be used for mRBS. Ten to 13 electrodes whose poststimulus histograms (PSTHs; summed across all electrodes) showed a peak larger than baseline activity were chosen for further use. A single stimulus amplitude was selected for all electrodes and for the entire duration of the measurement. Once the electrodes had been chosen, the experiment consisted of evaluation periods interleaved with short periods of mRBS (Fig. 1). Each evaluation period included 1), a sequence in which a subset of electrodes were probed such that PSTHs could be calculated; 2), a period in which there was only spontaneous activity; and 3) another probing sequence. Probing sequences involved all electrodes that were also used for mRBS in fully randomized order, using fixed amplitudes at 5 s intervals and repeated 15 times.

mRBS stimulation

Stimuli were generated in a probabilistic manner in both time and space. At every stepdt, one randomly selected electrode was allowed to be acti-vated. The probability of activation (i.e., the rate) was time-varying, as:

rðtÞ ¼ r0

þ r1

cosðumtÞ:

(1)

Where the angular velocity of modulation, um¼ 2pfm, was chosen such that

fm¼ 4 Hz. Random activation events with instantaneous probability r(t)

were generated by comparing at each time step a pseudo-random number x, uniformly distributed between 0 and 1, to r(t) $ dt. The resulting time series was Poisson-distributed with a time-varying parameter (i.e.,r(t)). The average rate of stimulation,r0, was set to 10 s1, which was enough

to significantly suppress bursting (as discussed in Results; but see also Wagenaar et al. (7)). The maximal deviation (r1) was set to either 8 or 10 s1.

The time stepdt was set as small as technically possible, which was deter-mined at 20 ms. The sequence of stimulation sites (10–13 sites) was fully randomized. Consecutive stimuli on the same electrode were allowed.

Phase-locked tetani

At particular phases of the modulation frequency of mRBS, tetani consisting of four pulses at a rate of 200 s1were applied to one electrode that was not already being used for RBS. All experiments consisted of at least three mRBS stimulation sessions with phase-locked tetani. The control group was formed by sham stimulation sessions in which no stimuli at all were applied. The parameters used are summarized inTable 1.

We use cosine notation, such that 0corresponds to peak mRBS stimula-tion rate (18 or 20 s1) and 180to the minimal mRBS rate of 0 or 2 s1(see

Fig. 1). In addition, because of setup limitations, mRBS was suspended from 40 ms before tetanus to 40 ms after tetanus. The number of pulses per tetanus was chosen in accordance with Huerta and Lisman (22), who found that four pulses (at 200 s1) applied at the peak or trough of the hippocampal q oscil-lation were enough to induce LTP or LTD (respectively) with the same magnitude as traditional protocols. Subsequent groups used four to five pulses at either 200 or 400 s1(400 s1chosen to limit tetanus duration) (21,23).

Probe all electrodes

Probe session Spontaneous activity m RBS stim ulation 5 1 0 6 6 30 Time [min] 0° sham 180° P1 P2 P3 P4

FIGURE 1 Timeline of a measurement. First, all electrodes were probed with stimuli of three different amplitudes. Based on the array-wide PSTHs, a selection of 10–13 electrodes was made for further use. Next, evaluation periods were interleaved with mRBS stimulation or sham periods of no stim-ulation. An evaluation period consisted of two probe sessions and 30 min of spontaneous recording. Both probe sessions were used, such that P1 and P2 were compared with P3 and P4 in the analysis. At least two different settings were used (0and 180;long arrows in insets indicate tetani) in random order.

(3)

PSTH analysis

The network state was evaluated with the use of PSTHs made during eval-uation periods. Electrode-specific PSTHs were made of responses to probe stimuli delivered through each electrode used also for mRBS. A PSTH for each electrode was calculated as the average histogram of activity at 0–500 ms after probe stimuli, using bins of 5 ms. The change in area of the PSTHs was taken as a measure for changes in network responses:

D

i;j

¼

area

after i;j

 area

before i;j

area

before i;j

(2)

Wherei is for every electrode, j denotes stimulus electrodes, and ‘‘before’’ and ‘‘after’’ are with respect to an mRBS intervention (Fig. 1). We disre-garded PSTHs with an areabeforesmaller than 10 spikes to decrease numeric

fluctuations caused by division by a small number. This also vastly reduced the amount of PSTHs that merely contained direct responses (i.e., responses that involved no or very few synapses). A normalized histogram of changes (all Di,jof all experiments) was made to visualize the aggregate results of

the applied interventions. We tested whether histograms originated from two different distributions by using a two-sample, single-sided Kolmo-gorov-Smirnov test (a¼ 0.05). Another measure that is more sensitive to the extremes of histograms is the potentiation ratio (PR), which is calculated as the number of potentiated (D > 20%) connections divided by the number of depressed connections (D < 20%). The PR takes into account the fact that most connections in the network remain unchanged. We derived the PR from the work of Chiappalone et al. (34), who also proposed the threshold of 20%. Although the relation between PSTHs (as described here) and synaptic plasticity is far from direct, several previous studies revealed concomitant changes when MEA activity was measured in parallel with intracellular activity (5,35–38).

Burst profile analysis

In addition to PSTHs, we analyzed bursts during spontaneous sessions using a previously described method (26). In brief, we calculated the array-wide spiking rate by convolving the spike-train with a Gaussian with a resolution of 1 ms. Whenever the resulting signal crossed a threshold, a part of the signal centered to the first local maximum after threshold crossing was selected. We also calculated the spiking rates of individual electrodes by convolving spike trains with a Gaussian. Thresholds were adjusted to the culture’s activity. The standard deviation used to calculate single electrode spiking rates was 5 ms. Selected episodes extended from 200 ms before, to 400 ms after the time of the maximum array-wide spiking rate. This was wide enough to capture the whole burst in all measurements. In accordance with earlier work, we call the Gaussian-smoothed, array-wide firing rate the ‘‘burst profile’’, and the electrode-specific smoothed firing rate the ‘‘phase profile’’. Before conduct-ing further analyses, we calculated the averages of the profiles over 5 min bins. This reduced the variability and computational load. We treated these averaged profiles as if they were individual profiles. Changes in profiles were calculated as the magnitude of the difference:

change

i;j

¼ kprofile

j

 profile

i

k

isj

(3)

wherei and j enumerate bursts in a measurement, and profilejis a 601-element

(i.e., 600 ms) vector describing either a burst profile or a phase profile. To assess changes, we created two groups. The between-group consists of changes between profiles that are separated in time by mRBS intervention, and the within-group contains changes between profiles that are not separated in time by mRBS. The within-group represents natural variability of profiles, whereas the between-group represents variability plus a possible effect of intervention. We used Student’st-tests to determine significance (a¼ 0.05). For clarity of presentation, the between-group changes depicted in Fig. 6 were normalized to the mean within-group changes.

RESULTS

mRBS

Fig. 2

gives an overview of mRBS combined with tetani

applied at 0



. The average rate of stimulation (r0

) was 10 s

1

,

modulated with a rhythm (fm

) of 4 Hz and deviation (r1

)

of 8 s

1

. We observed that activity lagged stimulation by

~13 ms. This corresponds to the typical response time of

the culture to stimuli, which is also reflected in the typical

PSTH. The interspike interval distribution of observed

activ-ity during mRBS strongly resembles that of the stimuli, with

a relative increase at intervals near 0.25 s due to 4 Hz

modu-lation. In contrast, the interspike intervals of spontaneous

activity are more broadly distributed, with the longer lags

(e.g., > 0.5 s) corresponding to the relatively quiet times

between bursts. It is particularly interesting that the number

of intervals smaller than 20 ms is ~5 times lower during

stim-ulation than during spontaneous activity, indicating less

bursting. The right panels in

Fig. 2

illustrate the burst

suppression by mRBS. By investigating rhythmic activity

instead of bursting activity, one can study plasticity in a

more controlled manner.

PSTHs

The PSTHs showed two phases in the responses (

Fig. 3

).

Many electrodes measured activity only at very short

laten-cies (<10 ms). Such early responses were dominated by

direct responses, which are antidromic or orthodromic action

potentials resulting from activation of an axon (

39

). Activity

at longer latencies involves synaptic transfers, and was

usually part of activity patterns that resembled network

bursts. In the majority of cases, the shapes of PSTHs before

and after intervention were similar, whereas the area

under-neath the curve changed.

In most experiments, a mix of increased and decreased

PSTH areas was found, even when one probe electrode

was considered, as shown in

Fig. 3

. The implication is that

changes were pathway-specific rather than stimulus-site

specific. Such an observation argues against the possibility

that the excitability of the stimulated neuron was somehow

changed. Rather, it suggests that the synaptic connections

between stimulated and observed neurons were changed.

Aggregate results of the changes in PSTH area of all

indi-vidual electrodes that were observed for the two different

mRBS settings and control experiments are shown in

TABLE 1 Parameters used for several modes of mRBS

stimulation

Intervention Phase Pulses ITI

No. of experiments No. of cultures mRBS (in phase) 0 4 5 s 20 8 mRBS (out of phase) 180 4 5 s 20 8 Control (no stimulation) - - - 13 8 A sham stimulation session of equal duration was taken as control. Tetani were repeated every 5 s during 6 min of mRBS (seeFig. 1). The last two columns of

(4)

Fig. 4

. When tetani were delivered in-phase (at 0



) of the

modulation frequency, the PSTH areas increased

signifi-cantly from those observed when no stimulation was applied

(p

¼ 0.047, single-sided, two-sample Kolmogorov-Smirnov

test). Conversely, tetani delivered in antiphase (at 180



)

did not result in a significant decrease in overall PSTH

area (p

¼ 0.061).

The overall trend was toward a decrease in PSTH area.

This is best seen by dividing the area of the curve showing

a >20% increase by the area of the curve showing

a >20% decrease. The PR was 0.62 when no stimulation

11

100 ms 2 sp/ bin

81

18

17

FIGURE 3 Example of PSTH changes. PSTHs in response to probe stimuli delivered to electrode 17 (dark shaded) are shown before (gray) and after (black) in-phase (0) mRBS stimulation. In the complete experi-ment, mRBS stimulation was applied through all shaded electrodes, and tet-ani were delivered through electrode 45 (hatched). Twenty-two PSTHs with an area over 10 (spikes) were considered responsive. Of these, the area of 13 PSTHs increased by at least 20%, whereas that of only two PSTHs decreased by at least 20%. The bin width used was 5 ms. Electrodes are numbered in column-row fashion. -100 -80 -60 -40 -20 0 20 40 60 80 100 0 0.5 1 1.5 2 In-phase mRBS Antiphase mRBS Control (sham)

Changes in PSTH area (%)

Fre

que

nc

y of observation (normalized)

FIGURE 4 Aggregate results for changes in PSTH area. Without stimula-tion, the dotted curve was found. The mean change in the absence of mRBS (D) was3.1%, indicating an overall decrease of PSTH area. With tetani applied in-phase (0) with mRBS, a shift toward increases in PSTH area was found (D:þ5.7%; solid line). The shaded area indicates the standard deviation of the histogram. Slight decreases in PSTH area were found for tetani applied at 180of mRBS (D:6.2%). Bin width: 2.5%.

0 10 20 30 40 50 60

0

40

80

120

160

Time [s]

Spikes/bin

Time [s]

Spikes/bin

0 10 20 30 40 50 60

0

40

80

120

160

1

3

2

0.01 0.1

1 10

0.01

0.1

1

10

100

0

0.2 0.4 0.6 0.8

1

0

10

20

30

40

50

60

70

80

90

Rate [/s]

Time [s]

Observations [normalized]

Inter event intervals [s]

1,

3

FIGURE 2 mRBS stimulation and activity. Top left panel: Histograms of stimuli (black) and activity (dashed) with tetani applied at 0. Activity lagged stimuli by ~13 ms and was scaled by a factor 0.2 for clarity. The activity was successfully modulated by 4 Hz mRBS stimulation. Lower left panel: Interevent interval distribution. Curves 1 and 2 show the interspike interval distributions during stimula-tion and spontaneous activity, respectively. Curve 3 shows the interstimulus interval distribution. Under stimulation, long quiescent periods disappeared, and intense short-latency firing decreased (by ~80%). Right panels: Congre-gate spike rate, in bins of 0.25 s, for 1 min of activity without stimulation (top panel) and during mRBS stimula-tion (lower panel). Stimulastimula-tion increased overall activity but also suppressed bursts and general variability in the overall spike rate.

(5)

was applied, 0.39 for tetani applied at 180



of the modulated

stimulation, and 1.84 for tetani applied at 0



.

Burst analysis

To examine the effect of our interventions on spontaneous

activity, we recorded 30 min of bursting activity before

and after each intervention (

Fig. 1

). The firing rates of

indi-vidual neurons during a burst for one particular example are

shown in

Fig. 5

. Although there is usually a short period in

which all contributing electrodes synchronize, there are also

differences between contributing electrodes. For example,

electrode 13 was usually early to fire, whereas electrode 86

contributed mainly to the late phase of bursts. The specificity

of electrode contributions (which we call phase profiles) is

caused by the underlying network connectivity and pathways

of activation. Changes in activity during bursts included

a shift in the time of the main phase of firing, and more (or

less) intense firing. In

Fig. 5

, the tendency was toward

broader profiles and higher peak activity, indicating that

the bursts had become more intense as a whole.

Overall, the analysis of bursts shows that in-phase tetani

were able to induce significant changes in activity patterns,

both in the aggregate firing rate (burst profiles,

p

¼ 0.040,

two-sided Student’s

t-test) and in individual electrode

contributions (phase profiles,

p

¼ 0.025). Tetani delivered

in antiphase of the mRBS were unable to induce changes

in either burst profiles (p

¼ 0.182) or phase profiles

(p

¼ 0.134). In fact, there was a tendency toward

stabiliza-tion of existing patterns, as the mean distance between

profiles was slightly smaller when antiphasic stimulation

was applied than in controls in which no stimulation was

applied.

Fig. 6

summarizes the statistical analysis of changes

in spontaneous bursts.

DISCUSSION

In this study, we investigated whether rhythmic activity in

culture could produce phase-dependent plasticity. To this

end, we first required that naturally occurring network bursts

be replaced by rhythmic activity. This was successfully

achieved by modulated and stochastic RBS, which decreased

the percentage of small, interspike intervals that would

nor-mally be found in network bursts, and at the same time

increased overall activity. These small interspike intervals

are most likely to induce a spike-timing-dependent synaptic

plasticity that may otherwise obscure or negate

experimen-tally induced changes in connectivity (

4

).

The PSTH analysis showed a general decrease in

respon-siveness for the no-stimulation controls, which is not

uncommon in regularly stimulated cultured neuronal

net-works (

40,41

). However, when tetani were applied at 0



,

an increase in overall PSTH area was observed. The decrease

in PSTH area by tetani applied at 180



was not significant,

but it may have been partially obscured by the general

decrease of PSTH area found in controls. Nevertheless,

applying mRBS with phase-locked tetani had an effect

in networks of dissociated neocortical neurons, which

depended on the phase at which the tetani were applied.

200 ms

50 /s

11

18

81

FIGURE 5 Examples of phase profiles before (gray) and after (black) mRBS stimulation with tetani at 0. The profiles are averages of the last 10 min of spontaneous recording before mRBS stimulation, and the first 10 min after. Spontaneous recordings from the same experiment as in

Figs. 3 and 4were used.

0.4 0.6 0.8 1.0 1.2 1.4 1.6

Change (normalized)

Control

In-phase mRBS

Antiphase mRBS

Changes in burst profiles

*

0.4 0.6 0.8 1.0 1.2 1.4 1.6

Change (normalized)

Control

In-phase mRBS

Antiphase mRBS

Changes in phase profiles

*

FIGURE 6 Changes in burst profiles (left panel) and phase profiles (right panel) for different settings of mRBS stimulation. The bars indicate the mean change after intervention, relative to inherent variability of profiles. The mean changes induced by mRBS stimulation are compared with ongoing changes in the absence of stimuli (control). All profiles from all experiments were included. Error bars denote standard error of mean; *p < 0.05, two-sided Student’s t-test.

(6)

The burst analysis showed results similar to those obtained

by PSTHs, with only tetani applied in-phase with mRBS

resulting in a significant increase in distances between

pro-files as compared to controls. We did not test the duration

of these changes, but they extended at least 1 h after

interven-tion. In this study, recordings of spontaneous activity started

15 min after intervention and were stopped 15 min before

the next intervention. Therefore, the results from the burst

analysis may have been negatively influenced, especially

when one considers that probe stimuli were delivered during

these two blocks of 15 min, which may alter activity patterns

by themselves (

31

).

There may be another explanation for the reduced effect of

synaptic depression. It is possible that the average synaptic

efficacy is relatively low, and thus there is less opportunity

for a further decrease. Since most studies have focused on

potentiation using extracellular recordings, there is no

evi-dence to support this. However, Jimbo et al. (

5,36

) confirmed

a PSTH decrease by whole-cell patch-clamp recordings,

demonstrating that depression is possible and that PSTHs

are sensitive to depression.

Overall, our results indicate that tetani applied in-phase

with mRBS had a definite effect, whereas the effect of

anti-phasic tetani was much smaller at best. This difference in

effect suggests the importance of the network-wide firing

rate at the moment the tetani are applied. The efficacy of

tet-ani may be time-varying because mRBS modulates the firing

threshold for neurons that are postsynaptic to the stimulus

site. Because stimuli were generated at a relatively low

average rate in a probabilistic way, and mRBS was

discon-tinued from 40 ms before tetanus to 40 ms afterward, it is

unlikely that such scaling of excitability is a direct result of

stimulation. Rather, mRBS may excite neurons to oscillate

at an intrinsic oscillation frequency near the modulation

fre-quency. Cortical neurons have a resonance frequency near

4 Hz (

42,43

). In this respect, it is interesting that Wagenaar

et al. (

9,10

) previously combined RBS (fixed aggregate

rate of 50 s

1

, cyclic electrode switching) with tetanic

stim-ulation (20 trains, 20 pulses/train, 20 s

1

, 2 s between trains)

but found no significant changes in responses to test stimuli.

Our study shows that rhythmic activity and the phase of

tetanus delivery are important factors that strongly influence

further processing in a neuronal network.

The authors thank Remy Wiertz for preparation and maintenance of the cultures.

REFERENCES

1. Baruchi, I., and E. Ben-Jacob. 2007. Towards neuro-memory-chip: imprinting multiple memories in cultured neural networks.Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 75:050901.

2. Gross, G. W., E. Rieske, ., A. Meyer. 1977. A new fixed-array multi-microelectrode system designed for long-term monitoring of extracel-lular single unit neuronal activity in vitro.Neurosci. Lett. 6:101–105. 3. Shahaf, G., and S. Marom. 2001. Learning in networks of cortical

neurons.J. Neurosci. 21:8782–8788.

4. le Feber, J., J. van Pelt, and W. L. Rutten. 2009. Latency-related devel-opment of functional connections in cultured cortical networks. Biophys. J. 96:3443–3450.

5. Jimbo, Y., H. P. Robinson, and A. Kawana. 1998. Strengthening of synchronized activity by tetanic stimulation in cortical cultures: applica-tion of planar electrode arrays.IEEE Trans. Biomed. Eng. 45:1297–1304. 6. Maeda, E., Y. Kuroda, ., A. Kawana. 1998. Modification of parallel activity elicited by propagating bursts in developing networks of rat cortical neurones.Eur. J. Neurosci. 10:488–496.

7. Wagenaar, D. A., R. Madhavan, ., S. M. Potter. 2005. Controlling bursting in cortical cultures with closed-loop multi-electrode stimula-tion.J. Neurosci. 25:680–688.

8. Wagenaar, D. A., Z. Nadasky, and S. M. Potter. 2006. Persistent dynamic attractors in activity patterns of cultured neuronal networks. Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 73, 051907.

9. Wagenaar, D. A., J. Pine, and S. M. Potter. 2006. Searching for plas-ticity in dissociated cortical cultures on multi-electrode arrays.J. Negat. Results Biomed. 5:16.

10. Wagenaar, D. A., J. Pine, and S. M. Potter. 2007. Correction: searching for plasticity in dissociated cortical cultures on multi-electrode arrays. J. Negat. Results Biomed. 6:3.

11. Ben-Ari, Y. 2001. Developing networks play a similar melody.Trends Neurosci. 24:353–360.

12. Buzsa´ki, G., and A. Draguhn. 2004. Neuronal oscillations in cortical networks.Science. 304:1926–1929.

13. Leinekugel, X., R. Khazipov, ., G. Buzsa´ki. 2002. Correlated bursts of activity in the neonatal hippocampus in vivo.Science. 296:2049–2052. 14. Meister, M., R. O. Wong, ., C. J. Shatz. 1991. Synchronous bursts of action potentials in ganglion cells of the developing mammalian retina. Science. 252:939–943.

15. Corner, M. A., J. van Pelt, ., R. H. Nuytinck. 2002. Physiological effects of sustained blockade of excitatory synaptic transmission on spontaneously active developing neuronal networks—an inquiry into the reciprocal linkage between intrinsic biorhythms and neuroplasticity in early ontogeny.Neurosci. Biobehav. Rev. 26:127–185.

16. Dravid, S. M., and T. F. Murray. 2004. Spontaneous synchronized calcium oscillations in neocortical neurons in the presence of physiolog-ical [Mg(2þ)]: involvement of AMPA/kainate and metabotropic gluta-mate receptors.Brain Res. 1006:8–17.

17. Bakkum, D. J., Z. C. Chao, and S. M. Potter. 2008. Spatio-temporal electrical stimuli shape behavior of an embodied cortical network in a goal-directed learning task.J. Neural Eng. 5:310–323.

18. Chao, Z. C., D. J. Bakkum, and S. M. Potter. 2008. Shaping embodied neural networks for adaptive goal-directed behavior.PLOS Comput. Biol. 4:e1000042.

19. Chao, Z. C., D. J. Bakkum, D. A. Wagenaar, and S. M. Potter. 2005. Effects of random external background stimulation on network synaptic stability after tetanization.Neuroinformatics. 3:263–280.

20. Geinisman, Y. 2000. Structural synaptic modifications associated with hippocampal LTP and behavioral learning.Cereb. Cortex. 10:952–962. 21. Ho¨lscher, C., R. Anwyl, and M. J. Rowan. 1997. Stimulation on the positive phase of hippocampal q rhythm induces long-term potentiation that can be depotentiated by stimulation on the negative phase in area CA1 in vivo.J. Neurosci. 17:6470–6477.

22. Huerta, P. T., and J. E. Lisman. 1995. Bidirectional synaptic plasticity induced by a single burst during cholinergic q oscillation in CA1 in vitro.Neuron. 15:1053–1063.

23. Hyman, J. M., B. P. Wyble, ., M. E. Hasselmo. 2003. Stimulation in hippocampal region CA1 in behaving rats yields long-term potentiation when delivered to the peak of q and long-term depression when deliv-ered to the trough.J. Neurosci. 23:11725–11731.

24. Paulsen, O., and T. J. Sejnowski. 2000. Natural patterns of activity and long-term synaptic plasticity.Curr. Opin. Neurobiol. 10:172–179. 25. Hasselmo, M. E., C. Bodelo´n, and B. P. Wyble. 2002. A proposed

func-tion for hippocampal q rhythm: separate phases of encoding and retrieval enhance reversal of prior learning.Neural Comput. 14:793–817.

(7)

26. Stegenga, J., J. Le Feber, ., W. L. Rutten. 2008. Analysis of cultured neuronal networks using intraburst firing characteristics.IEEE Trans. Biomed. Eng. 55:1382–1390.

27. Van Pelt, J., M. A. Corner, ., G. J. Ramakers. 2004. Longterm stability and developmental changes in spontaneous network burst firing patterns in dissociated rat cerebral cortex cell cultures on multielectrode arrays. Neurosci. Lett. 361:86–89.

28. van Pelt, J., P. S. Wolters, ., G. J. Ramakers. 2004. Long-term char-acterization of firing dynamics of spontaneous bursts in cultured neural networks.IEEE Trans. Biomed. Eng. 51:2051–2062.

29. Madhavan, R., Z. C. Chao, and S. M. Potter. 2007. Plasticity of recur-ring spatiotemporal activity patterns in cortical networks.Phys. Biol. 4:181–193.

30. Stegenga, J., J. Le Feber, ., W. L. Rutten. 2009. The effect of learning on bursting.IEEE Trans. Biomed. Eng. 56:1220–1227.

31. Vajda, I., J. van Pelt, ., A. van Ooyen. 2008. Low-frequency stimula-tion induces stable transistimula-tions in stereotypical activity in cortical networks.Biophys. J. 94:5028–5039.

32. Romijn, H. J., F. van Huizen, and P. S. Wolters. 1984. Towards an improved serum-free, chemically defined medium for long-term culturing of cerebral cortex tissue.Neurosci. Biobehav. Rev. 8:301–334. 33. Wagenaar, D. A., T. B. DeMarse, and S. M. Potter. 2005. MeaBench: a toolset for multi-electrode data acquisition and online analysis. Proc. Int. IEEE EMBS Conf. Neural Eng., 2nd, Arlington, VA. 34. Chiappalone, M., P. Massobrio, and S. Martinoia. 2008. Network

plas-ticity in cortical assemblies.Eur. J. Neurosci. 28:221–237.

35. Jimbo, Y., A. Kawana, ., V. Torre. 2000. The dynamics of a neuronal culture of dissociated cortical neurons of neonatal rats.Biol. Cybern. 83:1–20.

36. Jimbo, Y., T. Tateno, and H. P. C. Robinson. 1999. Simultaneous induc-tion of pathway-specific potentiainduc-tion and depression in networks of cortical neurons.Biophys. J. 76:670–678.

37. Maeda, E., H. P. Robinson, and A. Kawana. 1995. The mechanisms of generation and propagation of synchronized bursting in developing networks of cortical neurons.J. Neurosci. 15:6834–6845.

38. Pancrazio, J. J., E. W. Keefer, ., G. W. Gross. 2001. Neurophysiologic effects of chemical agent hydrolysis products on cortical neurons in vitro.Neurotoxicology. 22:393–400.

39. Bakkum, D. J., Z. C. Chao, and S. M. Potter. 2008. Long-term activity-dependent plasticity of action potential propagation delay and amplitude in cortical networks.PLoS One. 3:e2088.

40. Eytan, D., N. Brenner, and S. Marom. 2003. Selective adaptation in networks of cortical neurons.J. Neurosci. 23:9349–9356.

41. Jimbo, Y., H. P. Robinson, and A. Kawana. 1993. Simultaneous measurement of intracellular calcium and electrical activity from patterned neural networks in culture. IEEE Trans. Biomed. Eng. 40:804–810.

42. Siapas, A. G., E. V. Lubenov, and M. A. Wilson. 2005. Prefrontal phase locking to hippocampal q oscillations.Neuron. 46:141–151. 43. Ulrich, D. 2002. Dendritic resonance in rat neocortical pyramidal cells.

Referenties

GERELATEERDE DOCUMENTEN

The theoretical equation derivation to determine the type of sequence network and the positive sequence impedance for an A-phase-to-earth fault is shown below. It is important to

constater nettement comment le mur de chaînage roman fut con- struit dans une large tranchée de fandation taillée dans le sol vierge, et remblayée ensuite avec

Door zijn positie, zijn afmetingen en zijn profiel kan deze gracht vergeleken worden met deze van de circulaire versterking te Oost- Souburg (Zeeland).. Nadere

Voor de behandeling als moderator van het verband tussen affectieve empathie en delinquentie kwam naar voren dat studies met respondenten die in behandeling zijn kleinere

Existing crisis research largely fails to address the effects of different social media on stakeholder perceptions in crisis situations (Schultz, Utz, &amp; Goritz,

bevestiging-vragende, niet-sociaal angstige deelnemers. Op basis van deze uitkomst wordt verwacht dat sociaal angstige mensen ook in Facebookposts meer om bevestiging vragen. Zo

Het kan zijn dat tijdens de oefening je aandacht afgeleid wordt, dit is niet erg, merk het op en probeer je aandacht weer terug naar je ademhaling te sturen.. Let erop dat je nog in

Studies concerning polymeric micelles include, but are not limited to, application in the fields of active target- ing anticancer drugs to tumors [ 25 ], imaging of various