circuit topology for synchronizing neurons in …neuronet.jp/pdf/o_096.pdfcircuit topology for...

13
Circuit topology for synchronizing neurons in spontaneously active networks Naoya Takahashi a , Takuya Sasaki a , Wataru Matsumoto a , Norio Matsuki a , and Yuji Ikegaya a,b,1 a Laboratory of Chemical Pharmacology, Graduate School of Pharmaceutical Sciences, University of Tokyo, Tokyo 113-0033, Japan; and b Precursory Research for Embryonic Science and Technology, Japan Science and Technology Agency, Kawaguchi 332-0012, Japan Edited by Robert Desimone, Massachusetts Institute of Technology, Cambridge, MA, and approved April 15, 2010 (received for review December 19, 2009) Spike synchronization underlies information processing and stor- age in the brain. But how can neurons synchronize in a noisy network? By exploiting a high-speed (5002,000 fps) multineuron imaging technique and a large-scale synapse mapping method, we directly compared spontaneous activity patterns and anatomical connectivity in hippocampal CA3 networks ex vivo. As compared to unconnected pairs, synaptically coupled neurons shared more common presynaptic neurons, received more correlated excitatory synaptic inputs, and emitted synchronized spikes with approxi- mately 10 7 times higher probability. Importantly, common presyn- aptic parents per se synchronized more than unshared upstream neurons. Consistent with this, dynamic-clamp stimulation revealed that common inputs alone could not account for the realistic de- gree of synchronization unless presynaptic spikes synchronized among common parents. On a macroscopic scale, network activity was coordinated by a power-law scaling of synchronization, which engaged varying sets of densely interwired (thus highly synchro- nized) neuron groups. Thus, locally coherent activity converges on specic cell assemblies, thereby yielding complex ensemble dy- namics. These segmentally synchronized pulse packets may serve as information modules that ow in associatively parallel net- work channels. action potential | calcium imaging | microcircuit | spontaneous activity | synchronization S ynchronized spikes prevail in cortical networks (1, 2). Their modulations are commonly found in relation to attention, sensory processing, and motor behaviors (38) and are implied in perceptual binding (9). From computational aspects, spike syn- chronization is crucial in information propagation (10, 11). As single synapses are weak and stochastic, spikes cannot propagate to a downstream network unless they synchronize. Moreover, synchronized spikes are known to induce long-lasting synaptic plasticity, depending on their relative timings between pre- synaptic and postsynaptic neurons (12, 13). Spikes do not synchronize only between a pair of neurons but also among a set of neurons, often yielding high-order complex dynamics, as in cell assemblies (1417), synre chains (10, 18), and neuronal avalanches (19). The complex dynamics usually emerge through autoassociative recurrent networks in which neurons are sparsely interconnected to constitute relatively small groups. Thus, a neural correlation can report a local network state (2022). Independent studies have addressed the topology underlying either functional(synchronous) or anatomical(synaptic) connectivity among multiple neurons (2334), but very little is known about their relationship. In this work, we used high-speed functional multineuron calcium imaging (fMCI), large-scale synapse mapping, and multiple whole-cell and dynamic patch- clamp recording techniques and directly compared spatiotem- poral spike patterns with synaptic wiring architectures. We re- port that CA3 networks ex vivo are nonrandomly woven to facilitate local spike synchronization under globally coherent inhibitory backgrounds. Results Strong Spike Synchronization Between Synaptically Connected Neurons. In rat entorhino-hippocampal slice cultures, 104 pairs of adjacent CA3 pyramidal cells (PCs) were randomly selected for whole-cell recordings (Fig. 1A). The connectivity density among CA3 PCs was 28.8%. This connection ratio is higher than the connectivity (28%) reported in acute slice preparations (35, 36). In acute hippocampal slices, 75% to 90% of the axons of CA3 pyramidal neurons are amputated even in 500-μm-thick slices (37). Therefore, organotypically cultured ex vivo networks are likely to self-restore their complexity to a realistic extent. In support of this, we found that neither levels nor patterns of spontaneous excitatory postsynaptic currents (sEPSCs) or in- hibitory postsynaptic currents (sIPSCs) differed between ex vivo and in vivo hippocampal neurons (Fig. S1). As the ex vivo re- covery of slice cultures occurs without external inputs (unlike normal development), this work will describe the defaultnet- work dynamics that emerge under disturbance-free conditions. Of 104 PC pairs, we encountered 16 bidirectionally connected pairs. Given the connection probability p and the total number of pairs N, the statistically expected number of bidirectional pairs would be Np 2 , i.e., 8.6 pairs. In our datasets, therefore, the number of bidirectionally connected pairs was 1.9 times higher than expected (P = 0.012) (30), suggesting that CA3 recurrent networks are topologically biased to enhance local connectivity. When the neuron pairs were held in current-clamp mode, we sometimes found spontaneous spikes synchronized (Fig. 1B). To quantify the synchrony level, we introduce the scalar measure S, based on statistical salience of the observed number of syn- chronized spikes relative to chance. Specically, we counted all synchronized spikes, i.e., any pairs of spikes that concurred in two neurons within a given time lag. If neuron i and neuron j are independent units that re in a random manner, the probability P(n) that they exhibit n synchronized spikes during the obser- vation period t is given by the Poisson equation: P i; j ðnÞ¼ m n i; j n! e mi: j [1] where m i,j is the expected number of synchronized spikes, i.e., f i × f j × t, and f i and f j denote the spike rates of neuron i and neuron j , respectively. When spikes synchronize n times, the probability P i;j (rareness) is given as follows: P i; j ¼ k¼n P i; j ðkÞ¼ 1 −∑ n 1 k¼0 P i; j k [2] Then, we dened the surprise index (S i,j ) as log 2 P i; j (38). Author contributions: N.T., N.M., and Y.I. designed research; N.T. and T.S. performed research; N.T. and W.M. analyzed data; and N.T. and Y.I. wrote the paper. The authors declare no conict of interest. 1 To whom correspondence should be addressed. E-mail: [email protected]. This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10. 1073/pnas.0914594107/-/DCSupplemental. 1024410249 | PNAS | June 1, 2010 | vol. 107 | no. 22 www.pnas.org/cgi/doi/10.1073/pnas.0914594107

Upload: others

Post on 22-Aug-2020

7 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Circuit topology for synchronizing neurons in …neuronet.jp/pdf/O_096.pdfCircuit topology for synchronizing neurons in spontaneously active networks Naoya Takahashia, Takuya Sasakia,

Circuit topology for synchronizing neurons inspontaneously active networksNaoya Takahashia, Takuya Sasakia, Wataru Matsumotoa, Norio Matsukia, and Yuji Ikegayaa,b,1

aLaboratory of Chemical Pharmacology, Graduate School of Pharmaceutical Sciences, University of Tokyo, Tokyo 113-0033, Japan; and bPrecursory Researchfor Embryonic Science and Technology, Japan Science and Technology Agency, Kawaguchi 332-0012, Japan

Edited by Robert Desimone, Massachusetts Institute of Technology, Cambridge, MA, and approved April 15, 2010 (received for review December 19, 2009)

Spike synchronization underlies information processing and stor-age in the brain. But how can neurons synchronize in a noisynetwork? By exploiting a high-speed (500–2,000 fps) multineuronimaging technique and a large-scale synapse mapping method, wedirectly compared spontaneous activity patterns and anatomicalconnectivity in hippocampal CA3 networks ex vivo. As comparedto unconnected pairs, synaptically coupled neurons shared morecommon presynaptic neurons, received more correlated excitatorysynaptic inputs, and emitted synchronized spikes with approxi-mately 107 times higher probability. Importantly, common presyn-aptic parents per se synchronized more than unshared upstreamneurons. Consistent with this, dynamic-clamp stimulation revealedthat common inputs alone could not account for the realistic de-gree of synchronization unless presynaptic spikes synchronizedamong common parents. On a macroscopic scale, network activitywas coordinated by a power-law scaling of synchronization, whichengaged varying sets of densely interwired (thus highly synchro-nized) neuron groups. Thus, locally coherent activity converges onspecific cell assemblies, thereby yielding complex ensemble dy-namics. These segmentally synchronized pulse packets may serveas information modules that flow in associatively parallel net-work channels.

action potential | calcium imaging | microcircuit | spontaneous activity |synchronization

Synchronized spikes prevail in cortical networks (1, 2). Theirmodulations are commonly found in relation to attention,

sensory processing, and motor behaviors (3–8) and are implied inperceptual binding (9). From computational aspects, spike syn-chronization is crucial in information propagation (10, 11). Assingle synapses are weak and stochastic, spikes cannot propagateto a downstream network unless they synchronize. Moreover,synchronized spikes are known to induce long-lasting synapticplasticity, depending on their relative timings between pre-synaptic and postsynaptic neurons (12, 13).Spikes do not synchronize only between a pair of neurons but

also among a set of neurons, often yielding high-order complexdynamics, as in cell assemblies (14–17), synfire chains (10, 18), andneuronal avalanches (19). The complex dynamics usually emergethrough autoassociative recurrent networks in which neurons aresparsely interconnected to constitute relatively small groups. Thus,a neural correlation can report a local network state (20–22).Independent studies have addressed the topology underlying

either “functional” (synchronous) or “anatomical” (synaptic)connectivity among multiple neurons (23–34), but very little isknown about their relationship. In this work, we used high-speedfunctional multineuron calcium imaging (fMCI), large-scalesynapse mapping, and multiple whole-cell and dynamic patch-clamp recording techniques and directly compared spatiotem-poral spike patterns with synaptic wiring architectures. We re-port that CA3 networks ex vivo are nonrandomly woven tofacilitate local spike synchronization under globally coherentinhibitory backgrounds.

ResultsStrong Spike Synchronization Between Synaptically ConnectedNeurons. In rat entorhino-hippocampal slice cultures, 104 pairsof adjacent CA3 pyramidal cells (PCs) were randomly selectedfor whole-cell recordings (Fig. 1A). The connectivity densityamong CA3 PCs was 28.8%. This connection ratio is higher thanthe connectivity (2–8%) reported in acute slice preparations (35,36). In acute hippocampal slices, 75% to 90% of the axons ofCA3 pyramidal neurons are amputated even in 500-μm-thickslices (37). Therefore, organotypically cultured ex vivo networksare likely to self-restore their complexity to a realistic extent. Insupport of this, we found that neither levels nor patterns ofspontaneous excitatory postsynaptic currents (sEPSCs) or in-hibitory postsynaptic currents (sIPSCs) differed between ex vivoand in vivo hippocampal neurons (Fig. S1). As the ex vivo re-covery of slice cultures occurs without external inputs (unlikenormal development), this work will describe the “default” net-work dynamics that emerge under disturbance-free conditions.Of 104 PC pairs, we encountered 16 bidirectionally connected

pairs. Given the connection probability p and the total number ofpairs N, the statistically expected number of bidirectional pairswould be Np2, i.e., 8.6 pairs. In our datasets, therefore, thenumber of bidirectionally connected pairs was 1.9 times higherthan expected (P = 0.012) (30), suggesting that CA3 recurrentnetworks are topologically biased to enhance local connectivity.When the neuron pairs were held in current-clamp mode, we

sometimes found spontaneous spikes synchronized (Fig. 1B). Toquantify the synchrony level, we introduce the scalar measure S,based on statistical salience of the observed number of syn-chronized spikes relative to chance. Specifically, we counted allsynchronized spikes, i.e., any pairs of spikes that concurred intwo neurons within a given time lag. If neuroni and neuronj areindependent units that fire in a random manner, the probabilityP(n) that they exhibit n synchronized spikes during the obser-vation period t is given by the Poisson equation:

Pi; jðnÞ ¼mn

i; j

n!e−mi: j [1]

where mi,j is the expected number of synchronized spikes, i.e.,fi × fj × t, and fi and fj denote the spike rates of neuroni andneuronj, respectively. When spikes synchronize n times, theprobability �Pi;j (rareness) is given as follows:

�Pi; j ¼ ∑∞

k¼nPi; jðkÞ ¼ 1− ∑

n− 1

k¼0Pi; j

�k�

[2]

Then, we defined the surprise index (Si,j) as− log2�Pi; j (38).

Author contributions: N.T., N.M., and Y.I. designed research; N.T. and T.S. performedresearch; N.T. and W.M. analyzed data; and N.T. and Y.I. wrote the paper.

The authors declare no conflict of interest.1To whom correspondence should be addressed. E-mail: [email protected].

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.0914594107/-/DCSupplemental.

10244–10249 | PNAS | June 1, 2010 | vol. 107 | no. 22 www.pnas.org/cgi/doi/10.1073/pnas.0914594107

Page 2: Circuit topology for synchronizing neurons in …neuronet.jp/pdf/O_096.pdfCircuit topology for synchronizing neurons in spontaneously active networks Naoya Takahashia, Takuya Sasakia,

For extremely synchronized pairs with S > 100 bits, S isdenoted as 100 bits to avoid arithmetic precision problems incomputing floating-point numbers. Thus, S ranges between 0 and100 bits, with higher S reporting stronger synchronization.S varies as a function of time widows for synchrony detection.

It peaked at time windows of 10 to 50 ms (Fig. 1C). In accordwith this, cross-correlation histogram of spike timings in neuronpairs (Fig. 1C Inset) showed a peak at 0 ms with an SD of 25 ms(fitted to the Gaussian curve). We therefore selected the 10-msbin in the following analysis, unless otherwise specified.Synaptically connected pairs were more synchronized than

uncoupled pairs (Fig. 1D); the mean S for synaptic pairs washigher by 24 bits than that for uncoupled pairs, indicating that,on average, synaptic pairs exhibited synchronized spikes witha 1.7 × 107 (i.e., 224) times higher probability. Consistent withthis, the probability to find synaptic pairs increased with S (Fig.S2). Therefore, the functional and anatomical connectivity isreciprocally, if not perfectly, linked at the single-cell level.

Common Presynaptic Neurons Shared by Synaptically ConnectedNeurons. To probe the synaptic wiring pattern among CA3 PCs,we employed reverse optical trawling (ROTing), a new opticalmapping method (39). While EPSCs were recorded simulta-neously from two randomly selected PCs, a small number ofnearby neurons were sequentially activated by iontophoreticapplication of glutamate through a glass pipette that was slowlymoved to survey the surrounding network (Fig. 1E). Spikes ofglutamate-activated neurons were captured as somatic Ca2+

transients with high-speed fMCI (Fig. S3). Neurons that exhibi-ted calcium transients immediately before EPSCs were statisti-cally screened. With this procedure, we can identify 96% ofpresynaptic cells projecting to the patch-clamped neurons (39).Fig. 1 E–G shows an example of ROTing-identified synapticconnections, where nine neurons were sequentially whole-cellrecorded in pair, and 66 neurons were imaged with fMCI.

In these synapse connection maps, we often found “common”presynaptic neurons that projected to both of the two patchedneurons. The mean ratio of common presynaptic neurons to thetotal presynaptic neurons that projected to at least one of thepatched neuron pair was 13.2± 2.5%. This proportion increased to43.9 ± 20.2%, however, when the postsynaptic pairs were synap-tically connected (Fig. 1H; n = 11 pairs). This suggests that syn-aptic pairs receive more correlated inputs than nonsynaptic pairs.To confirm this, we carried out targeted patch-clamp record-

ings from spontaneously synchronized PC pairs, which wereidentified online by high-speed fMCI. After monitoring sponta-neous spikes in current-clamp mode, sEPSCs and sIPSCs wererecorded in voltage-clamp mode at −90 mV and 0 mV, re-spectively. As expected, these PC pairs received highly correlatedsEPSCs and sIPSCs (Fig. 2A Left). The cross-correlations peakedsharply at time lags of less than 1 ms (Fig. 2A Right). The halfpeak width in the cross-correlogram was 67.0 ± 56.1 ms forsEPSCs and 54.7 ± 50.4 ms for sIPSCs.As control experiments, we also targeted less synchronized

neuron pairs. In the low-S pairs, sEPSCs were only weakly cor-related (Fig. 2B), whereas sIPSCs were still highly correlated(Fig. 2B). Data are summarized in Fig. 2 C and D. The corre-lation coefficients of sEPSCs correlated positively with S (r =0.34, P < 0.01), whereas those of sIPSCs were always high, in-dependent of S. These data imply that excitatory neurons arespecifically wired to ensure local synchronization and that in-hibitory activity is globally coherent in CA3 networks.

Synchronized Inputs from Common Presynaptic Neurons. How effi-ciently do common excitatory synaptic inputs produce synchro-nous firing? With the dynamic-clamp technique, we injectedartificially generated “correlated” synaptic conductance patternsinto twoCA3PCs under pharmacological blockade of fast synaptictransmission. The stimulus sweeps were constructed in silico fromspike trains of 200 Poisson-firing presynaptic neurons, in whicheach spike was convolved with a unitary waveform of a fast excit-

Fig. 1. Linkage of spike synchronization and synaptic connectivity. (A) Biocytin reconstruction of highly synchronized PCs. (B) Spontaneous activity exhibitedby the neuron pairs shown in A. Dots indicate synchronized spike pairs. (C) (Upper) Cross-correlation histogram showing the spike-triggered distribution ofspikes emitted by another neuron (n = 56 pairs). (Lower) S peaked at time windows of 10 to 50 ms (n = 39 pairs). (D) Synaptically connected pairs displayedsignificantly higher S than unconnected pairs (P = 0.032, Kolmogorov-Smirnov test). Data of unidirectional and bidirectional pairs were mixed, as there was nostatistical difference (P > 0.1). (E) Diagram of ROTing for synapse mapping. (Upper) Two postsynaptic neurons were voltage-clamped (blue), and theirpresynaptic neurons were searched based on timings between EPSCs in the patched neurons and glutamate-evoked presynaptic calcium events (red). (Lower)Representative synaptic responses found by ROTing. Postsynaptic synaptic currents were aligned at presynaptic calcium spike timings (60-ms jitters allowed;gray). Presynaptic cell 38 projected to postsynaptic cell 1, whereas cell 39 projected to cells 1 and 2. The locations of these cells are shown in F. (F) Repre-sentative synaptic connectivity, in which 63 connections were identified to project to nine patched neurons (color-filled circles). Color lines indicate synapticlinks to the correspondingly colored neurons. (G) Cells in E are circularly arranged. Arrows outside the circle indicate unidirectional (uni) or bidirectional (bi)synaptic connections between simultaneously patched neurons. (H) Synaptically coupled neuron pairs were innervated by larger proportions of commonpresynaptic neurons, compared with unconnected pairs (**P = 0.0043, Mann-Whitney U test; n = 5 unconnected pair vs. eight connected pairs).

Takahashi et al. PNAS | June 1, 2010 | vol. 107 | no. 22 | 10245

NEU

ROSC

IENCE

Page 3: Circuit topology for synchronizing neurons in …neuronet.jp/pdf/O_096.pdfCircuit topology for synchronizing neurons in spontaneously active networks Naoya Takahashia, Takuya Sasakia,

atory postsynaptic conductance. Correlated conductances weregenerated by overlapping some of the 200 presynaptic spike trainsso that their correlation coefficients ranged from 0 to 1 (Fig. 2E).When PC pairs were stimulated with these conductances, theyemitted spikes, some of whichwere found to synchronize (Fig. 2F).Swas calculated from the spike series and plotted against the inputcorrelation coefficients. The neurons responded to more corre-lated inputs with more synchronized spike outputs (Fig. 2G). Theinput-output relationship was substantially weak, however, com-pared with that observed during spontaneous activity (cf. Fig. 2C);note that S was always less than 50 bits in the stimulation experi-ments, whereas it often exceeded 50 bits even for weakly corre-lated EPSCs in spontaneous activity.This inconsistency between artificial and spontaneous synaptic

inputs suggests that naturally occurring EPSCs involve moreinformation than captured by simple cross-correlations, therebyefficiently synchronizing postsynaptic neurons (40, 41). To seekthe underlying information, we examined the dynamics ofspontaneous firing activity of CA3 neuron populations. Usinghigh-speed fMCI, the spatiotemporal pattern of spontaneousnetwork activity was reconstructed with the millisecond resolu-tion from 85 ± 25 neurons (n= 14 slices), ranging from 53 to 137neurons (Fig. 3A and Movie S1). Within a given 10-ms period, onaverage, only a small number of neurons (0.19 ± 0.61% of thetotal cells; n = 14 slices) were active, whereas the network oc-casionally exhibited large synchronous events that involved up to

approximately 50% neurons (Fig. 3A Lower). The frequency ofthe synchrony sizes was approximated by a power-law distribu-tion P(n) ≈ nα, where n denotes the synchrony size and P(n) theprobability of observing size-n synchrony (Fig. 3B; α = −2.6). Aperi-synchronization time histogram revealed that, in a synchro-nous event, the network barraged spikes during a period of ap-proximately ±20 ms (Fig. S4). During synchronization, CA1networks showed high-frequency field oscillations (Fig. S5),which closely resembled sharp-wave/ripples observed in thehippocampus of quiescent or sleeping animals (42).We computed S for all neuron pairs in a raster plot and drew

the S-based connectivity map (Fig. 3C). Synchronization wassparse; S was 0 bits in 80.3 ± 17.9% pairs, whereas nonzero Sconformed to a log-normal Gaussian distribution (Fig. 3D). S didnot correlate with the physical cell-to-cell distance as a whole (r=0.096, P > 0.1), but pairs with S > 50 bits were always locatedwithin a distance of 200 μm (Fig. 3E). Note that closely spacedspikes (>50 Hz firing) were hardly separated in fMCI data (Fig.S6A) because Ca2+ transients had a long decay constant of ap-proximately 500 ms (43). Therefore, fMCI tends to underestimateS, whereas S obtained by fMCI was almost linear with electro-physiologically obtained S (r = 0.77, P < 0.01; Fig. S6B).After constructing the S maps, we conducted ROTing in the

same network to search common presynaptic neurons. Thecommon presynaptic pairs exhibited significantly higher S thanunshared pairs (Fig. 3F; P < 0.01, Kolmogorov-Smirnov test).This suggests that common parents are more prone to synchro-nize. In other words, inputs into common postsynaptic pairs willbe more correlated than “Poisson-correlated” inputs, which weused in the dynamic-clamp experiments (Fig. 2E).Given that network synchronization magnitudes were power-

law tailed, we now generated the synaptic conductance patternsfrom power law-scaled presynaptic spike trains by keeping theoverall presynaptic firing rate the same as in the Poisson simu-lation (Fig. 3G). Neuron pairs stimulated by the scaled patternconductances exhibited strongly synchronized spikes, sometimesreaching an S of 100 bits (Fig. 3 H and I).

Ensemble Dynamics of Spontaneous Activity. Based on the matricesof S obtained by fMCI, we constructed dendrograms using themethod of Ward, a hierarchical clustering algorithm (Fig. S7A).It disclosed the “cliqueness” of synchronous neurons. We thenexamined whether the S matrices included small-world attributes(27). By setting various lower-limit thresholds on the S matrices,we depicted the synchrony graphs as a function of the threshold(Fig. S7B). For any threshold, the extracted graph exhibited the“small-worldness” (Fig. S7 C–E).Using the affinity propagation algorithm (44), the order of

neurons was sorted so that higher S pairs were more neighboredin the matrix, and the renumbered neurons were clustered intosubgroups. In the representative data shown in Fig. 4A, 96neurons were separated into 15 groups. On average, each moviecontained 16.1 ± 7.6 groups, and each group comprised 5.8 ± 4.1neurons (n = 14 slices). This clustering was validated withROTing; the connection probability among within-group neu-rons was 37.6 ± 17.8%, significantly higher than the across-groupconnectivity (27.0 ± 8.6%, P = 0.04, Wilcoxon signed-rank test;n = 5 slices; Fig. 4B).Neurons within a group were sparsely distributed over the

imaged field (Fig. 4C). The distribution of the cell-to-cell dis-tance between within-group neurons did not differ from that ofacross-group neurons (Fig. 4D; P > 0.1, Kolmogorov-Smirnovtest; n = 14 slices), indicating no spatial bias of synchronousspike patterns. Raster plots sorted along the groups demon-strated that within-group neurons frequently exhibited synchro-nized spikes (Fig. 4E Upper). Interestingly, the internal structureof synchronization was dynamic, that is, different synchronyevents recruited different sets of neuron groups (Fig. 4E Lower).

Fig. 2. Correlated sEPSCs in synchronous neuron pairs. (A) sEPSCs (Upper)and sIPSCs (Lower) recorded in a neuron pair with an S of 100. Blue and redtraces are parts of simultaneous voltage-clamp recordings from two pyra-midal neurons. (B) In a neuron pair with an S of 4.0, sEPSCs were only weaklyuncorrelated, but sIPSCs were still strongly correlated. (C and D) Summary ofthe relationship between S and the correlation coefficient of sEPSCs (C) (r =0.34, P = 0.01; n = 56 pairs) and sIPSCs (D) (r = 0.14, P > 0.1; n = 55 pairs). (E)With the dynamic-clamp technique, correlated excitatory synaptic con-ductances were injected into two CA3 pyramidal neurons under the phar-macological blockade of fast synaptic transmission. Two stimulation sweeps(red and blue) were generated from partially overlapped Poisson spiketrains. (F) Spike patterns emitted by two PCs in response to correlatedconductance stimuli (correlation coefficient of 0.6). (G) Swas calculated fromspike responses and plotted against the correlation coefficients between theinjected conductances (n = 55 pairs).

10246 | www.pnas.org/cgi/doi/10.1073/pnas.0914594107 Takahashi et al.

Page 4: Circuit topology for synchronizing neurons in …neuronet.jp/pdf/O_096.pdfCircuit topology for synchronizing neurons in spontaneously active networks Naoya Takahashia, Takuya Sasakia,

Information Flow from CA3 to CA1. Finally, we addressed how theseCA3 cell assembly activities are transferred to downstream CA1networks. After sorting CA3 neurons based on S matrices, we

applied ROTing to probe the CA3-to-CA1 synaptic connectivity.An example map is presented in Fig. 5A, which shows the pro-jection pattern from 90 CA3 neurons to seven patch-clamped

Fig. 3. Heavy-tailed synchronization in spontaneously active networks. (A) Rastergram of spontaneous spikes in 96 neurons monitored by high-speed fMCI(Upper). Time histogram of the percentage of coactivated cells to the total imaged neurons (Lower, 10-ms bin). (B) Power-law distributions in synchrony size(n = 14 slices). (C) Anatomical distribution of 96 neurons imaged in A (Upper). To quantify the degree of pairwise synchronization, S was defined as a surpriseindex of the probability that the observed number of synchronized spikes could occur under the Poisson process. Gray-scaled S between all pairs of 96 cellswas shown on the cell map (Lower). (D) Distribution of S (54,239 cell pairs from 14 slices). (E) Relationship between S and the physical distance between twoneurons (r = 0.096, P > 0.1). (F) Pairs of common presynaptic cells shared by any postsynaptic pairs exhibited higher S than the others (P < 0.01, Kolmogorov-Smirnov test). (G) Stimulation sweeps were generated from power law–distributed spike trains. (H) Spike patterns in response to power law–scaled con-ductance stimuli (correlation coefficient of 0.6). (I) S was plotted against the correlation coefficients of the scaled conductances (n = 66 pairs).

Fig. 4. Cell assembly dynamics in network synchrony. (A) Affinity propagation algorithm separated 96 neurons in Fig. 3A into 15 subgroups. Each color indicatesa single group. (B) ROTing was conducted after neuronal group classification. Connection probability is significantly higher for within-group pairs than across-group pairs (P = 0.04, Wilcoxon signed-rank test; n = 5 slices). (C) Cell map of two representative groups found in A. (D) No significant difference in cumulativedistribution of the cell-to-cell distances betweenwithin-group pairs (black line) and across-group pairs (gray line; n = 14 slices). (E) Raster plot (Fig. 3A) was sortedin the order of the groups defined in A. Five synchrony events are time-expanded (Lower). Different events consisted of different group sets.

Takahashi et al. PNAS | June 1, 2010 | vol. 107 | no. 22 | 10247

NEU

ROSC

IENCE

Page 5: Circuit topology for synchronizing neurons in …neuronet.jp/pdf/O_096.pdfCircuit topology for synchronizing neurons in spontaneously active networks Naoya Takahashia, Takuya Sasakia,

CA1 PCs. CA3 PC pairs that projected convergently to the sameCA1 PCs had statistically higher S than the other CA3 PC pairs(Fig. 5B; P < 0.01, Kolmogorov-Smirnov test). Thus, synchro-nous CA3 activity tends to converge on the same CA1 neurons.

DiscussionIn this study, we used high-speed fMCI to capture the fine-scalestructure of spontaneous CA3 network activity and analyzed theorganization and generation of spike synchronization with the aidof large-scale optical mapping and dynamic-clamp techniques.We found that complex recruitments of highly synchronized cellassemblies constitute large-scale network synchronization andthat the assemblies emerge through presynaptic synchronization.

Origin of Synchronization. More synchronized neuron pairs weremore likely to be synaptically linked. We do not think, however,that this synaptic link is enough to synchronize these neurons,because a single synapse is tooweak to depolarize beyond the spikethreshold. Rather, correlated synaptic inputs from multiple com-mon presynaptic neurons (30–33, 45) are more plausibly causal ofspike synchronization. Two lines of evidence are in favor of thishypothesis. First, ROTing revealed that synaptically coupled CA3PC pairs shared numerically more presynaptic CA3 PCs. Second,double whole-cell recordings of neuron pairs revealed that moresynchronized pairs received more correlated sEPSCs.We also found that common parent PCs were strongly syn-

chronized. This indicates that these common parents are alsounder the further-upstream innervation by common “grandpar-ent” PCs. This suggests synchronous activity flow, that is, syn-chronous pulse packets flow across densely connected neurongroups in recurrent networks. This idea is supported by our dy-namic clamp data showing that common presynaptic neuronsneed to be synchronized to evoke a realistic level of postsynapticsynchronization, although common inputs from randomly spik-ing neurons can do neurons to some extent.In the dynamic clamp experiments, however, we did not con-

sider the dendritic properties. Dendrites exhibit nonlinear exci-tation through spatiotemporal input summation and dendriticspikes. It is also possible that synchronized spikes result fromnonlinear dendritic computation with convergent inputs of syn-chronous activity into specific dendritic branches. Research intothis possibility is under way in our laboratory with multiplepatch-clamp recordings targeting dendrites (46).

Cell Assemblies in Small-World Networks. Consistent with previousreports showing that the functional and anatomical connectivityamong individual neurons exhibits small-world architectures (26,27, 47, 48), the CA3 networks also included small-world topol-ogy. The small-world network is theoretically believed to allowfast information transfer with low wiring costs (49), the co-existence of information segregation and integration (50), andsynchronization (51). Neurons within a small-world cluster,classified by affinity propagation, were sparsely distributed inspace. In hippocampal place cell activity recorded by multipleunit electrodes in vivo, it is also reported that synchronous cli-ques are dispersed across the electrodes (17). Despite this ap-parent randomness of neuron locations, within-group neuronswere preferentially interconnected. This may emerge from tar-get-selection mechanisms, such as activity-dependent synapticplasticity. Moreover, synchronous CA3 neuron groups convergedonto the same CA1 neurons. Thus, the CA3-to-CA1 connectivityforms relatively independent routes that carry CA3 ensembleactivities to specific CA1 neuron subsets (Fig. S8). Synchronousmodules may serve as endogenous building blocks that embodythe diversity and complexity of information processing in asso-ciative and parallel networks.

Coherent Inhibitory Networks. In the rat neocortex, the axons ofinhibitory neurons highly arborize, twisting and turning, seem-ingly rummaging among their postsynaptic targets (52). There-fore, single interneurons may promiscuously provide nearby PCswith correlated inhibitory inputs. Moreover, interneurons arereciprocally connected through synaptic contacts and gap junc-tion (53–55). Thus, their interplay may give rise to globally co-herent inhibition of PC networks (56). Indeed we found thatsIPSCs were correlated between virtually all PC pairs. Inhibitoryinputs limit the window available for temporal summation andincrease the temporal precision of PC firing (57). They alsoentrain network activity by interacting with intrinsic oscillatorymechanisms of PCs (58). We thus speculate that interneuronscouple relatively independent PC subgroups and orchestratecomplex network synchronization.

Ex Vivo Data.Our data were obtained exclusively from organotypicslice cultures, an ex vivo experimental network model, and hencemust be carefully extrapolated to other neuronal systems, suchmature brain networks in vivo. Nonetheless, it is still intriguing tofind that such spontaneously reorganized ex vivo networks showhighly nonrandom patterns of connectivity and activity. Ourfindings thus describe a primary regime of how neuronal networksintrinsically develop and operate in the “free-run” mode.

Materials and MethodsExperiments were performed with the approval of the animal experimentethics committee at the University of Tokyo (approval number 19-43, A21-6)according to the University of Tokyo guidelines for the care and use oflaboratory animals.

Entorhino-hippocampal organotypic slices were prepared from 7-d-oldWistar/ST rats (SLC). Experiments were performed on days 7 to 11 in vitro.Patch-clamp recordings were carried out simultaneously from two to four PCswith two Axopatch 700B dual amplifiers (Molecular Devices). Whole-cellpatch pipettes (4–6 MΩ) were filled with 135 mM K-gluconate, 4 mM KCl, 10mM Hepes, 10 mM phosphocreatine, 4 mM MgATP, 0.3 mM NaGTP, and0.2% biocytin. Dynamic-clamp stimulation was performed with a PCI-6024Edata acquisition board (National Instruments) under a real-time Linux-environment.

For fMCI, slices were incubated with Oregon Green 488 BAPTA-1 AM at37 °C for 1 h and imaged at 500 to 2,000 frames per s with a Nipkow-diskconfocal unit (CSUX-1; Yokogawa Electric), a high-speed back-illuminatedCCD camera (iXon DU860; Andor), and a water-immersion objective lens(magnification ×16, 0.80 NA; Nikon). Spike timings were determined with anautomatic machine-learning algorithm. Spike train data used here areavailable online at http://hippocampus.jp/data. Further details are providedin SI Materials and Methods.

Fig. 5. Convergent projection from synchronized CA3 pairs to CA1 neurons.(A) Representative wiring patterns between 91 CA3 and 7 CA1 neurons. Forillustration purpose, CA3 neurons were classified on their S matrix with theWard method. (B) CA3 neuron pairs that projected convergently to the sameCA1 neurons displayed significantly higher S than the others (P < 0.01,Kolmogorov-Smirnov test; n = 4 slices).

10248 | www.pnas.org/cgi/doi/10.1073/pnas.0914594107 Takahashi et al.

Page 6: Circuit topology for synchronizing neurons in …neuronet.jp/pdf/O_096.pdfCircuit topology for synchronizing neurons in spontaneously active networks Naoya Takahashia, Takuya Sasakia,

ACKNOWLEDGMENTS. We are grateful to Ms. Huei-Yu Chiou (GraduateSchool of Pharmaceutical Sciences, University of Tokyo) for providing us within vivo whole-cell recording data. This work was supported in part bya Grant-in-Aid for Science Research on Priority Areas (Elucidation of neural

network function in the brain, 18021008, 17023015, and 20019014); a Grant-in-Aid for Science Research (17650090, 17689004, and 19659013) from theMinistry of Education, Culture, Sports, Science, and Technology of Japan;and Sumitomo Foundation Grant 050038.

1. Engel AK, Fries P, Singer W (2001) Dynamic predictions: Oscillations and synchrony intop-down processing. Nat Rev Neurosci 2:704–716.

2. Salinas E, Sejnowski TJ (2001) Correlated neuronal activity and the flow of neuralinformation. Nat Rev Neurosci 2:539–550.

3. Zohary E, Shadlen MN, Newsome WT (1994) Correlated neuronal discharge rate andits implications for psychophysical performance. Nature 370:140–143.

4. Vaadia E, et al. (1995) Dynamics of neuronal interactions in monkey cortex in relationto behavioural events. Nature 373:515–518.

5. deCharms RC, Merzenich MM (1996) Primary cortical representation of sounds by thecoordination of action-potential timing. Nature 381:610–613.

6. Riehle A, Grün S, Diesmann M, Aertsen A (1997) Spike synchronization and ratemodulation differentially involved in motor cortical function. Science 278:1950–1953.

7. Stopfer M, Bhagavan S, Smith BH, Laurent G (1997) Impaired odour discrimination ondesynchronization of odour-encoding neural assemblies. Nature 390:70–74.

8. Steinmetz PN, et al. (2000) Attention modulates synchronized neuronal firing inprimate somatosensory cortex. Nature 404:187–190.

9. Gray CM, Singer W (1989) Stimulus-specific neuronal oscillations in orientationcolumns of cat visual cortex. Proc Natl Acad Sci USA 86:1698–1702.

10. Abeles M (1991) Corticonics: Neural circuits of the cerebral cortex (Cambridge UnivPress, Cambridge, UK).

11. Diesmann M, Gewaltig MO, Aertsen A (1999) Stable propagation of synchronousspiking in cortical neural networks. Nature 402:529–533.

12. Markram H, Lübke J, Frotscher M, Sakmann B (1997) Regulation of synaptic efficacyby coincidence of postsynaptic APs and EPSPs. Science 275:213–215.

13. Bi GQ, Poo MM (1998) Synaptic modifications in cultured hippocampal neurons:dependence on spike timing, synaptic strength, and postsynaptic cell type. J Neurosci18:10464–10472.

14. Hebb DO (1949) The organization of behavior: A neuropsychological theory (JohnWiley and Sons, New York).

15. Nicolelis MA, Baccala LA, Lin RC, Chapin JK (1995) Sensorimotor encoding bysynchronous neural ensemble activity at multiple levels of the somatosensory system.Science 268:1353–1358.

16. Laubach M, Wessberg J, Nicolelis MA (2000) Cortical ensemble activity increasinglypredicts behaviour outcomes during learning of a motor task. Nature 405:567–571.

17. Harris KD, Csicsvari J, Hirase H, Dragoi G, Buzsáki G (2003) Organization of cellassemblies in the hippocampus. Nature 424:552–556.

18. Ikegaya Y, et al. (2004) Synfire chains and cortical songs: temporal modules of corticalactivity. Science 304:559–564.

19. Beggs JM, Plenz D (2003) Neuronal avalanches in neocortical circuits. J Neurosci 23:11167–11177.

20. Schneidman E, Berry MJ, 2nd, Segev R, Bialek W (2006) Weak pairwise correlationsimply strongly correlated network states in a neural population. Nature 440:1007–1012.

21. Shlens J, et al. (2006) The structure of multi-neuron firing patterns in primate retina. JNeurosci 26:8254–8266.

22. Tang A, et al. (2008) A maximum entropy model applied to spatial and temporalcorrelations from cortical networks in vitro. J Neurosci 28:505–518.

23. Beggs JM, Plenz D (2004) Neuronal avalanches are diverse and precise activitypatterns that are stable for many hours in cortical slice cultures. J Neurosci 24:5216–5229.

24. Eytan D, Marom S (2006) Dynamics and effective topology underlying synchronizationin networks of cortical neurons. J Neurosci 26:8465–8476.

25. Sasaki T, Matsuki N, Ikegaya Y (2007) Metastability of active CA3 networks. J Neurosci27:517–528.

26. Yu S, Huang D, Singer W, Nikolic D (2008) A small world of neuronal synchrony. CerebCortex 18:2891–2901.

27. Watts DJ, Strogatz SH (1998) Collective dynamics of ‘small-world’ networks. Nature393:440–442.

28. Holmgren C, Harkany T, Svennenfors B, Zilberter Y (2003) Pyramidal cell communicationwithin local networks in layer 2/3 of rat neocortex. J Physiol 551:139–153.

29. Shepherd GMG, Stepanyants A, Bureau I, Chklovskii DB, Svoboda K (2005) Geometricand functional organization of cortical circuits. Nat Neurosci 8:782–790.

30. Song S, Sjöström PJ, Reigl M, Nelson S, Chklovskii DB (2005) Highly nonrandomfeatures of synaptic connectivity in local cortical circuits. PLoS Biol 3:e68.

31. Yoshimura Y, Dantzker JL, Callaway EM (2005) Excitatory cortical neurons form fine-scale functional networks. Nature 433:868–873.

32. Kampa BM, Letzkus JJ, Stuart GJ (2006) Cortical feed-forward networks for bindingdifferent streams of sensory information. Nat Neurosci 9:1472–1473.

33. Otsuka T, Kawaguchi Y (2008) Firing-pattern-dependent specificity of corticalexcitatory feed-forward subnetworks. J Neurosci 28:11186–11195.

34. Lefort S, Tomm C, Floyd Sarria JC, Petersen CC (2009) The excitatory neuronal networkof the C2 barrel column in mouse primar somatosensory cortex. Neuron 61:301–316.

35. Miles R, Wong RK (1986) Excitatory synaptic interactions between CA3 neurones inthe guinea-pig hippocampus. J Physiol 373:397–418.

36. Smith KL, Szarowski DH, Turner JN, Swann JW (1995) Diverse neuronal populationsmediate local circuit excitation in area CA3 of developing hippocampus. JNeurophysiol 74:650–672.

37. Gomez-Di Cesare CM, Smith KL, Rice FL, Swann JW (1997) Axonal remodeling duringpostnatal maturation of CA3 hippocampal pyramidal neurons. J Comp Neurol 384:165–180.

38. Weaver W (1948) Probability, rarity, interest, and surprise. Sci Mon 67:390–392.39. Sasaki T, Minamisawa G, Takahashi N, Matsuki N, Ikegaya Y (2009) Reverse optical

trawling for synaptic connections in situ. J Neurophysiol 102:636–643.40. Stevens CF, Zador AM (1998) Input synchrony and the irregular firing of cortical

neurons. Nat Neurosci 1:210–217.41. DeWeese MR, Zador AM (2006) Non-Gaussian membrane potential dynamics imply

sparse, synchronous activity in auditory cortex. J Neurosci 26:12206–12218.42. Buzsáki G (1989) Two-stage model of memory trace formation: A role for “noisy”

brain states. Neuroscience 3:551–570.43. Sasaki T, Takahashi N, Matsuki N, Ikegaya Y (2008) Fast and accurate detection of

action potentials from somatic calcium fluctuations. J Neurophysiol 100:1668–1676.44. Frey BJ, Dueck D (2007) Clustering by passing messages between data points. Science

315:972–976.45. Kazama H, Wilson RI (2009) Origins of correlated activity in an olfactory circuit. Nat

Neurosci 12:1136–1144.46. Ishikawa D, et al. (2010) Fluorescent pipettes for optically targeted patch-clamp

recordings. Neural Netw, in press.47. Bettencourt LM, Stephens GJ, Ham MI, Gross GW (2007) Functional structure of

cortical neuronal networks grown in vitro. Phys Rev E Stat Nonlin Soft Matter Phys 75:021915.

48. Pajevic S, Plenz D (2009) Efficient network reconstruction from dynamical cascadesidentifies small-world topology of neuronal avalanches. PLOS Comput Biol 5:e1000271.

49. Buzsáki G, Geisler C, Henze DA, Wang XJ (2004) Interneuron Diversity series: Circuitcomplexity and axon wiring economy of cortical interneurons. Trends Neurosci 27:186–193.

50. Sporns O, Zwi JD (2004) The small world of the cerebral cortex. Neuroinformatics 2:145–162.

51. Barahona M, Pecora LM (2002) Synchronization in small-world systems. Phys Rev Lett89:054101.

52. Stepanyants A, Tamás G, Chklovskii DB (2004) Class-specific features of neuronalwiring. Neuron 43:251–259.

53. Fukuda T, Kosaka T (2000) Gap junctions linking the dendritic network of GABAergicinterneurons in the hippocampus. J Neurosci 20:1519–1528.

54. Tamás G, Buhl EH, Lörincz A, Somogyi P (2000) Proximally targeted GABAergicsynapses and gap junctions synchronize cortical interneurons. Nat Neurosci 3:366–371.

55. Bartos M, et al. (2002) Fast synaptic inhibition promotes synchronized gammaoscillations in hippocampal interneuron networks. Proc Natl Acad Sci USA 99:13222–13227.

56. Poo C, Isaacson JS (2009) Odor representations in olfactory cortex: “Sparse” coding,global inhibition, and oscillations. Neuron 62:850–861.

57. Pouille F, Scanziani M (2001) Enforcement of temporal fidelity in pyramidal cells bysomatic feed-forward inhibition. Science 293:1159–1163.

58. Cobb SR, Buhl EH, Halasy K, Paulsen O, Somogyi P (1995) Synchronization of neuronalactivity in hippocampus by individual GABAergic interneurons. Nature 378:75–78.

Takahashi et al. PNAS | June 1, 2010 | vol. 107 | no. 22 | 10249

NEU

ROSC

IENCE

Page 7: Circuit topology for synchronizing neurons in …neuronet.jp/pdf/O_096.pdfCircuit topology for synchronizing neurons in spontaneously active networks Naoya Takahashia, Takuya Sasakia,

Supporting InformationTakahashi et al. 10.1073/pnas.0914594107SI Materials and MethodsAnimal Experiment Ethics. Experiments were performed with theapproval of the animal experiment ethics committee at theUniversity of Tokyo (approval number 19-43, A21-6) according tothe University of Tokyo guidelines for the care and use of lab-oratory animals.

Slice Culture Preparations. Entorhino-hippocampal organotypicslices were prepared from 7-day-old Wistar/ST rats (SLC) as pre-viously described (1). Briefly, rat pups were anesthetized by hy-pothermia and decapitated. The brains were removed and placedin aerated, ice-cold Gey balanced salt solution supplementedwith 25 mM glucose. Horizontal entorhino-hippocampal sliceswere made at a thickness of 300 μm by a vibratome (DTK-1500;Dosaka). They were placed on Omnipore membrane filters(JHWP02500; Millipore) and incubated in 5% CO2 at 37 °C. Theculture medium, composed of 50% MEM (Invitrogen), 25%Hanks balanced salt solution, 25% horse serum (Cell CultureLaboratory), and antibiotics, was changed every 3.5 days. Ex-periments were performed on days 7 to 11 in vitro. Althoughslice cultures are known to self-rewire and form abnormal con-nections that very rarely exist in normal conditions, such as CA1-to-CA1, CA1-to-CA3, and CA3–to–dentate gyrus connections(2, 3), these aberrant connections are not dominant in our sliceculture preparations. ROTing, a synapse mapping techniquedescribed later, demonstrates that these abnormal connectionsare less than 0.5% of the total connections and that an over-whelming number of the connections project to their normaltargets. This is probably because in our preparations, the en-torhinal cortex is not dissected from slices of the hippocampalformation. Lesions of the entorhinal cortex are known to resultin abnormal sprouting and reorganization of hippocampal net-works in vivo and ex vivo (4, 5).

Ex Vivo Patch-Clamp Recordings. A slice was placed in a recordingchamber perfused at 3 to 4mL/minwith artificial cerebrospinalfluid(aCSF), consisting of 127mMNaCl, 26mMNaHCO3, 3.3mMKCl,1.24 mM KH2PO4, 1.0 mM MgSO4, 1.0 to 1.2 mM CaCl2, and10 mM glucose at 30 to 32 °C. Whole-cell recordings were carriedout simultaneously from two to four pyramidal cells. Patch pipettes(4–6MΩ) were filled with 135mMK-gluconate, 4mMKCl, 10mMHepes, 10 mM phosphocreatine, 4 mMMgATP, 0.3 mM NaGTP,and 0.2% biocytin. Single units were extracellularly recorded in theloose-cell–attached mode with aCSF-filled pipettes. To examinewhether the recorded neurons were synaptically connected, aCSFwas modified to 2.2 mM K+, 3.0 mM Mg2+, and 3.6 mM Ca2+ toreduce spontaneous activity and enhance the reliability of synaptictransmission (6).

In Vivo Patch-Clamp Recordings. Male ICR mice (18–20 day old)were anesthetized with urethane (1.2–1.8 g/kg, i.p.). Animals wereimplanted with a metal head-holder and mounted on a custom-made stereotaxic fixture. A small craniotomy (approximately1 mm2) was made at 2.5 mm caudal to the bregma and 2.2 mmventrolateral to the sagittal suture along the surface of the skull,and the dura was removed. Whole-cell recordings were obtainedwith the blind patch-clamp approach (7).

Data Acquisition. All electrophysiological recordings were carriedout using MultiClamp 700B amplifiers (Molecular Devices).Signals were low-pass filtered at 2 kHz and digitized at 20 kHz.Data were analyzed with custom-written scripts in IgorPro 6

(Wavemetrics). Excitatory and inhibitory postsynaptic current(EPSC and IPSC, respectively) were recorded at clamped vol-tages of –90 mV and 0 mV, respectively. EPSGs and IPSGs werecomputed on the assumption that EPSCs were recorded at thereversal potential for inhibition and that IPSCs were recorded atthe reversal potential for excitation.

Dynamic-Clamp Stimulation. CA3 PCs were stimulated with thedynamic-clamp conductance injection technique. Synaptic eventswere modeled based on conductance g(t), and the commandcurrent signal I(t) was computed as a function of I(t) = g(t) ×[V(t) – Erev] (V(t), membrane potential; Erev, reversal potential)under a real-time Linux environment and delivered into patch-clamped neurons at 20 kHz by a PCI-6024E data acquisitionboard (National Instruments). Conductance stimuli (30 s) con-sisted of a series of excitatory synaptic inputs, which were con-structed from 200 Poisson spike trains convolved with a unitaryconductance transient representing a fast excitatory synaptic re-sponse fitted by the dual exponential function (g0 × [exp(–t/τd) –exp(–t/τr)]), where τr represents an activation time constant,τd a decay time constant, and g0 a scaling factor (τr=0.5ms, τd=2ms, and g0 = 1,000 pS). The average rate of the total input firingwas 800 Hz. Two partially correlated conductance sweeps weresimultaneously generated so that their correlation coefficient Cinranged from 0 to 1 at roughly every 0.2 step; for an expectedcorrelation level (c), two sweeps, g1(t) and g2(t), were constructedby overlapping (1 – c) × 200 Poisson trains. To make the scaledsynaptic input sweep, presynaptic spikes were distributed ina power law across trains. During stimulation, intrinsic fast syn-aptic transmission was blocked in the presence of an inhibitormixture of 20 μM 6-cyano-7-nitroquinoxoxaline-2,3-dione, 50 μMD,L-2-amino-5-phosphonopentanoic acid, and 50 μM picrotoxin.

fMCI. Slices were incubated with 2 mL dye solution at 37 °C for 1 h(8). The dye solution was aCSF containing 0.0005% OregonGreen 488 BAPTA-1 (OGB-1) AM, 0.01% Pluronic F-127, and0.005% Cremophor EL. After 1 h recovery, a slice was trans-ferred to a recording chamber. Images were acquired at 500 to2,000 frames per s with a Nipkow-disk confocal unit (CSUX-1;Yokogawa Electric), a high-speed back-illuminated CCD camera(iXon DU860; Andor), a water-immersion objective lens (mag-nification ×16, 0.80 NA; Nikon), and Solis software (Andor).Fluorophores were excited at 488 nm with an argon laser (10–15mW, 532-BS-AO4; Omnichrome) and visualized with a 507-nmlong-pass emission filter. In each cell body, the fluorescencechange ΔF/F was calculated as (Ft – F0) / F0, where Ft is thefluorescence intensity at frame time t, and F0 is baseline. Spiketimings were determined as the onsets of individual Ca2+ tran-sients with an automatic machine-learning algorithm that canaccurately detect the timings within one-frame-jitter errors (9).

Local Field Potential Recordings and Ripple Detection. In someexperiments, CA1 local field potentials were recorded duringfMCI monitoring of the calcium activity of CA3 neurons. Glasspipettes were filled with 2 M NaCl and placed in CA1 stratumpyramidale. To extract the ripple wave activity, the recorded datawere band-pass filtered at 150 to 300 Hz. Ripple-like events wereautomatically detected based on their oscillatory powers anddurations; the root mean square (3-ms window) of the band-passed signal was used to detect the ripple wave with a powerthreshold of 5 SDs with 10 ms in duration.

Takahashi et al. www.pnas.org/cgi/content/short/0914594107 1 of 7

Page 8: Circuit topology for synchronizing neurons in …neuronet.jp/pdf/O_096.pdfCircuit topology for synchronizing neurons in spontaneously active networks Naoya Takahashia, Takuya Sasakia,

ROTing. The ROTing technique was used to find synapticallycoupled neurons located in an fMCI-imaged region (6). Imme-diately after monitoring spontaneous activity with fMCI, aCSFwas modified to 2.2 mM K+, 3.0 mM Mg2+, and 3.6 mM Ca2+ toreduce spontaneous activity and the resultant plasticity of syn-aptic wiring that may occur. In slices loaded with OGB-1 AM,two CA3 or CA1 PCs were voltage-clamped at –70 mV, and 10μM glutamate was locally puffed through iontophoretic pipettes(approximately 1 MΩ, 3–10 μA for each 1–5 s) to evoke spikes ina few CA3 neurons located in a fMCI-targeted region. The pi-pette was slowly moved over CA3 networks, and the evokedspikes were monitored by fMCI at 50 frames per s with an AndorDV897 CCD camera. These spike timings were statisticallycompared with EPSCs recorded in the patch-clamped neurons todetermine the responsible presynaptic pyramidal neurons.

Nissl and Biocytin Staining. After each experiment, the slice wasfixed in 4% paraformaldehyde overnight at 4 °C and 0.2% TritonX-100 overnight and then incubated overnight at 4 °C withNeuroTrace (N-21482, 1:100; Molecular Probes) or streptavidin–Alexa Fluor 594 conjugate (1:1,000; Invitrogen-MolecularProbes) for Nissl or biocytin staining, respectively. Based on bi-ocytin staining, data obtained from PCs were selected to analyzepatch-clamp data.

Surprise Index as Pairwise Correlation. We focused on pairwisespike correlations (10, 11), rather than higher-order correlations,because this study aimed to attribute spike synchronization tosynaptic connection, i.e., a structural relationship between twoneurons. Our high-speed scanning did not allow imaging for

a period of more than 3 min, and the numbers of spontaneousspikes were often insufficient to precisely calculate the correla-tion coefficient (12). To estimate the pairwise similarity in ourpoint-process dataset, therefore, we considered the probabilitythat spikes can synchronize by chance. Synchronized spike pairs(SSPs), defined as any pairs of spikes that concurred in twoneurons, were detected with a time window of 10 ms. If neuroniand neuronj are independent units that fire in a random manner,the probability P(n) that they exhibit n SSPs during the obser-vation period t is given by the Poisson equation:

Pi;jðnÞ ¼mn

i; j

n!e−mi:j

where mi,j is the expected number of SSPs, i.e., fi × fj × t, and fiand fj denote the spike rates of neuroni and neuronj, respectively.When SSPs occur n times, the probability (i.e., rareness) is:

�Pi;j ¼ ∑∞

k¼nPi; jðkÞ ¼ 1− ∑

n− 1

k¼0Pi; jðkÞ:

The surprise index (Si,j) was defined as follows (13):

− log2�Pi; j:

For extremely synchronized pairs with an S of more than 100 bits,S is denoted as 100 bits to avoid arithmetic precision problems incomputing floating-point numbers.

Data Representation.We reported all averaged values as means ±SDs.

1. Koyama R, et al. (2007) A low-cost method for brain slice cultures. J Pharmacol Sci 104:191–194.

2. De Simoni A, Griesinger CB, Edwards FA (2003) Development of rat CA1 neurones inacute versus organotypic slices: Role of experience in synaptic morphology andactivity. J Physiol 550:135–147.

3. Gähwiler BH, Capogna M, Debanne D, McKinney RA, Thompson SM (1997)Organotypic slice cultures: A technique has come of age. Trends Neurosci 20:471–477.

4. Laurberg S, Zimmer J (1981) Lesion-induced sprouting of hippocampal mossy fiber col-laterals to the fascia dentata in developing and adult rats. J Comp Neurol 200:433–459.

5. West JR, Dewey SL (1984) Mossy fiber sprouting in the fascia dentata after unilateralentorhinal lesions: Quantitative analysis using computer-assisted image processing.Neuroscience 13:377–384.

6. Sasaki T, Minamisawa G, Takahashi N, Matsuki N, Ikegaya Y (2009) Reverse opticaltrawling for synaptic connections in situ. J Neurophysiol 102:636–643.

7. MargrieTW,BrechtM,SakmannB (2002) In vivo, low-resistance,whole-cell recordings fromneurons in the anaesthetized and awake mammalian brain. Pflugers Arch 444:491–498.

8. Takahashi N, Sasaki T, Usami A, Matsuki N, Ikegaya Y (2007) Watching neuronalcircuit dynamics through functional multineuron calcium imaging (fMCI). NeurosciRes 58:219–225.

9. Sasaki T, Takahashi N, Matsuki N, Ikegaya Y (2008) Fast and accurate detection ofaction potentials from somatic calcium fluctuations. J Neurophysiol 100:1668–1676.

10. Schneidman E, Berry MJ, 2nd, Segev R, Bialek W (2006) Weak pairwise correlationsimply strongly correlated network states in a neural population. Nature 440:1007–1012.

11. Shlens J, et al. (2006) The structure of multi-neuron firing patterns in primate retina. JNeurosci 26:8254–8266.

12. de la Rocha J, Doiron B, Shea-Brown E, Josić K, Reyes A (2007) Correlation betweenneural spike trains increases with firing rate. Nature 448:802–806.

13. Weaver W (1948) Probability, rarity, interest, and surprise. Sci Mon 67:390–392.14. Humphries MD, Gurney K, Prescott TJ (2006) The brainstem reticular formation is

a small-world, not scale-free, network. Proc R Soc Lond B Biol Sci 273:503–511.

Takahashi et al. www.pnas.org/cgi/content/short/0914594107 2 of 7

Page 9: Circuit topology for synchronizing neurons in …neuronet.jp/pdf/O_096.pdfCircuit topology for synchronizing neurons in spontaneously active networks Naoya Takahashia, Takuya Sasakia,

Fig. S1. Spontaneous synaptic activity in hippocampal slice cultures is similar to that in in vivo hippocampus. (A) Representative whole-cell patch-clamp tracesof spontaneous excitatory postsynaptic current (Upper: sEPSC at –90 mV) and inhibitory postsynaptic current (Lower: sIPSC at 0 mV) recorded from hippocampalneurons in a slice culture preparation (ex vivo) and the hippocampus of a urethane-anesthetized mouse (in vivo). (B) We calculated the mean and SD as well asthe third and fourth standardized moments, i.e., skewness, and kurtosis, of spontaneous excitatory postsynaptic conductance (Upper: sEPSG) and inhibitorypostsynaptic conductance (Lower: sIPSG; n = 3 cells for each). No parameters did show statistical differences. Data are means ± SEM.

Fig. S2. The ratio of synaptic pairs increases as a function of S. CA3 PCs were randomly selected, and their spontaneous spike activities were recorded incurrent-clamp mode. The probability to find synaptically connected pairs were plotted against the spike synchronicity S.

Takahashi et al. www.pnas.org/cgi/content/short/0914594107 3 of 7

Page 10: Circuit topology for synchronizing neurons in …neuronet.jp/pdf/O_096.pdfCircuit topology for synchronizing neurons in spontaneously active networks Naoya Takahashia, Takuya Sasakia,

Fig. S3. High-speed fMCI. (A) Confocal images of the CA3 PC layer in an OGB-1 AM–loaded (Left) and post hoc Nissl–stained (Center) slice. Neurons are Nissl-positive and thus distinguishable from nonneuronal cells. (B) Simultaneous loose-patch recording and calcium imaging reveals that action potentials evokesomatic calcium transients. (C) Twenty neurons were monitored at 2,000 frames per s. Spontaneous ΔF/F traces of individual neurons (Right), the locations ofwhich are also shown (Left).

Fig. S4. Temporally sparse synchronization in spontaneously spiking CA3 networks. (A) (Upper) Rastergram of spontaneous spikes in 96 neurons monitored byfMCI (same as Fig. 3A in the main text). (Lower) Time histogram of the percentage of coactivated cells to the total imaged neurons (2-ms bin). (B) Time expansionof a single synchrony event marked by an asterisk inA. (C) Peri-synchronization time histograms in which the peak time of each synchrony event (≥15 × SDs fromthe mean activity) was aligned at time 0 ms (bin, 2 ms; n = 985 events of 14 slices). (D) Power-law distributions in synchrony size for various bin sizes (2, 10, 20, 50,and 100 ms) with the scaling exponents of −3.4, −2.6, −2.3, −2.0, and −1.9, respectively (n = 14 slices). The power law was robust for bin sizes of 2 to 100 ms.

Takahashi et al. www.pnas.org/cgi/content/short/0914594107 4 of 7

Page 11: Circuit topology for synchronizing neurons in …neuronet.jp/pdf/O_096.pdfCircuit topology for synchronizing neurons in spontaneously active networks Naoya Takahashia, Takuya Sasakia,

Fig. S5. Synchronization is accompanied by ripple-like high-frequency oscillations. Local field potentials (band-filtered at 150–300 Hz) were recorded fromCA1 stratum radiatum (Upper), whereas CA3 network activity was monitored with fMCI (Lower). The event marked with an asterisk was magnified in time(Right). Ripple-like high-frequency oscillations occurred with CA3 network synchronization.

Fig. S6. Comparison of optically and electrophysiologically determined S values. (A) Distribution of interspike intervals in spontaneous activity of neuronsrecorded by whole-cell current-clamp technique (n = 154 cells; black) and high-speed fMCI (n = 1,193 cells; green). The latter often missed “bursty” spikes withintervals of less than 20 ms. (B) Relationship between S of intracellularly recorded neuron pairs and its modified value (S’) in which spikes in greater than 50-Hzburst firings, except for each first spike, were removed from the spike series. Each dot represents a single neuron pair. The gray line indicates the linear di-agonal S = S’, and the green the linear least-square fit.

Takahashi et al. www.pnas.org/cgi/content/short/0914594107 5 of 7

Page 12: Circuit topology for synchronizing neurons in …neuronet.jp/pdf/O_096.pdfCircuit topology for synchronizing neurons in spontaneously active networks Naoya Takahashia, Takuya Sasakia,

Fig. S7. Small-world architectures in synchrony-based connectivity. (A) The S matrix shown in Fig. 3A in the main text is hierarchically shown in a dendrogramformat. (B) In the data for Fig. 3A in the main text, synchronized pairs were extracted at the S thresholds of 25, 15, and 5 (bits). Circles and lines in each graphindicate the extracted neurons and suprathreshold links, respectively. (C and D) Two metrics, i.e., the mean clustering coefficient C (a measure of how fre-quently neighbors of each cell are also neighbors of each other) and the mean shortest path length L (the average length of the shortest path between twocells), were compared with those of 200 equivalent random graphs. The random graphs maintained the total numbers of neurons and links, but all links wererandomly rewired. Chance is represented as the mean (line) ± SD (shade) of 200 randomly rewired graphs. At any given S threshold, C was consistently largerthan chance (Crand), whereas Lwas close to chance (Lrand). (E) For a small-world network, the ratios λ = L/Lrand and γ = C/Crand are expected to be approximately 1and greater than 1, respectively. Therefore, we calculated the scalar measure of small-worldness, defined as γ/λ (14). The small-worldness was consistentlyhigher than the chance level, i.e., 1, over the entire S threshold range.

Takahashi et al. www.pnas.org/cgi/content/short/0914594107 6 of 7

Page 13: Circuit topology for synchronizing neurons in …neuronet.jp/pdf/O_096.pdfCircuit topology for synchronizing neurons in spontaneously active networks Naoya Takahashia, Takuya Sasakia,

Fig. S8. Network organization and cell assemblies. The connectivity of CA3 PCs is densely clustered among specific neuron subsets (different colors). Theseinterwired neurons thus receive correlated excitatory inputs and thereby generate highly synchronized spikes. CA3 inhibitory interneurons provide stronglycorrelated background inhibition, which may integrate CA3 PC groups. The synchronized CA3 spikes are specifically transmitted to subsets of CA1 PCs throughrelatively independent channels (arrows).

Movie S1. Spontaneous activity of CA3 neurons in a slice loaded with OGB-1 AM. Time is compressed by a factor of 2.

Movie S1

Takahashi et al. www.pnas.org/cgi/content/short/0914594107 7 of 7