You are on page 1of 12

a r t ic l e s

Flexible information routing by transient synchrony


Agostina Palmigiano1–4, Theo Geisel1–3, Fred Wolf1–4,6 & Demian Battaglia2,5,6   
Perception, cognition and behavior rely on flexible communication between microcircuits in distinct cortical regions.
The mechanisms underlying rapid information rerouting between such microcircuits are still unknown. It has been proposed
that changing patterns of coherence between local gamma rhythms support flexible information rerouting. The stochastic and
transient nature of gamma oscillations in vivo, however, is hard to reconcile with such a function. Here we show that models
© 2017 Nature America, Inc., part of Springer Nature. All rights reserved.

of cortical circuits near the onset of oscillatory synchrony selectively route input signals despite the short duration of gamma
bursts and the irregularity of neuronal firing. In canonical multiarea circuits, we find that gamma bursts spontaneously arise
with matched timing and frequency and that they organize information flow by large-scale routing states. Specific self-organized
routing states can be induced by minor modulations of background activity.

Perception and cognition rely on context-dependent selection of rel- Here we investigate a class of circuit models that naturally exhibits
evant inputs and flexible interareal brain communication. However, extensive power, frequency and timing variability and that contains key
despite its fundamental role, the basis of dynamic information routing types of heterogeneity, including heterogeneous transmission delays.
remains an unsolved problem. In particular, the circuit mechanisms In these models, below the onset of developing oscillatory synchrony,
that underlie the rapid reconfiguration of selective information trans- collective oscillations are short-lived, with durations on the order
fer on behavioral time-scales are poorly understood. Proposals range of a few cycles, are weakly synchronized and exhibit stochastically
from hypothetical circuitry dedicated to routing1–3 to conditional sig- drifting frequencies. When multiple circuits with these characteris-
nal propagation4–6 to the hypothesis that neuronal oscillations direct tics are coupled by long-range excitatory connections, the resulting
interareal communication, either by enabling conditional readout large-scale dynamics spontaneously generates temporally co-occur-
through frequency filtering7 or by exploiting coherence in networks of ring bursts of synchrony. We find that the drifting frequencies of each
regular oscillations8. Windows for efficient communication along dif- region track each other, giving rise to transient phase-locking within
ferent pathways might potentially also arise from fluctuations of local the gamma bursts. Through state-resolved information-theoretical
network excitability. In particular, this latter ‘communication through analyses, we assessed whether these transient patterns of coherence
coherence’ hypothesis has been invoked in the interpretation of many can gate information flow. We further examined the propagation of
experimental findings in which boosted interareal coherence seems external input signals relayed from different source regions, as well as
associated with enhanced interareal communication9. For instance, under which conditions these signals can be decoded from the activity
increased interareal gamma-band coherence and phase synchroniza- of a downstream target region. We find that the transient coherence
tion might be markers of selective attention, in which different input between the activities of the local circuits dynamically shapes the flow
information streams have to be transmitted or attenuated according of information between them: information transfer is either selec-
to task relevance10,11. tively boosted or suppressed along different routes according to the
Conceptually, however, there are fundamental obstacles to a func- transient phase pattern. These distinctive routing states have a direc-
tion of neuronal oscillations in routing. First, oscillatory synchroni- tion set by the transient phase-relations and are modulated by the
zation is confined to short episodes lasting on the order of ~100 ms, fluctuating level of synchrony. We find that the propagation of exter-
and the statistical properties of these episodes bear strong signatures nally supplied information can be selectively gated on or off along
of stochasticity12,13. Second, neuronal spike firing is irregular and different pathways depending on the routing state. Unexpectedly, the
only weakly synchronized, with spikes emitted at every possible phase stochastic and fleeting nature of ongoing oscillations more effectively
within the ongoing oscillation cycle14. Third, interareal synaptic modulates information flow than stronger and more coherent forms
transmission delays are long and diverse and may counteract reli- of synchronized circuit activity.
able phase synchronization15. Fourth, the frequency of the transient
oscillatory bursts fluctuates over time, varies between recording sites RESULTS
and depends on characteristics of the presented stimuli16,17. A pri- The transient synchrony regime
ori, the likelihood of gamma bursts to spontaneously match should Oscillatory neuronal activity in vivo is comprised of epochs of syn-
thus be very low. chronous activity arising from an overall poorly synchronized state.

1Max Planck Institute for Dynamics and Self-organization, Göttingen, Germany. 2Bernstein Center for Computational Neuroscience, Göttingen, Germany. 3Institute for
Nonlinear Dynamics, Georg-August University School of Science, Göttingen, Germany. 4SFB-889 Cellular Mechanisms of Sensory Processing, Göttingen, Germany.
5Aix-Marseille Université, Inserm, Institut de Neurosciences des Systèmes, Marseille, France. 6These authors contributed equally to this paper. Correspondence

should be addressed to A.P. (agos@nld.ds.mpg.de) or D.B. (demian.battaglia@univ-amu.fr).

Received 22 November 2016; accepted 25 April 2017; published online 22 May 2017; doi:10.1038/nn.4569

nature NEUROSCIENCE  advance online publication 


a r t ic l e s

a b
No. 300

...
No. 1

16%

10 mV

100 ms

c Synchronization d Mean population e LFP spectrogram f Distribution of


index frequency (Hz) (a.u.) gamma bursts (a.u.)
10 3 10–3
40 40
2.5 2
0.8 80
2 2
50
© 2017 Nature America, Inc., part of Springer Nature. All rights reserved.

2 50

Frequency (Hz)
Frequency (Hz)
0.6 60
4 4
�in (kHz)
�in (kHz)

1.5
60 60 1
6 0.4 6 40
1
70 70
8 0.2 8 20 0.5

10 10 80 80 0
0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1 1 2 0 0.1 0.2 0.3 0.4
PI PI Time (s) Burst length (s)

Figure 1  The transient synchrony regime. (a) Scheme of the local connectivity of a brain region or area (inhibitory and excitatory neurons in blue and
red, respectively). (b) Simulated activity in the transient synchrony regime. Transient oscillatory bursts in LFP time-series (bottom) and multiunit activity
(middle, measured as percentage of active neurons). Individual neurons fire irregularly and not at every oscillation cycle, as indicated by the raster
plot of 300 excitatory neurons (top). (c) The degree of synchronization of local neuronal activity, as quantified by the synchronization index (Online
Methods), grows with the probability PI of recurrent inhibitory connections and the rate νin of the Poisson background drive to the region.
(d) Dependence of the frequency of collective oscillation on PI and νin (see Supplementary Fig. 1c for single-neuron firing rates). Dashed circles in c
and d highlight two working points, with similar collective oscillation frequencies in the gamma range but very different synchronization levels. Upper
(black) dashed circle (transient synchrony): PI = ~0.25 and νin = ~2.5 kHz. Central (gray) dashed circle (high synchrony): PI = ~0.55 and νin = ~5.5 kHz.
(e) Transient gamma bursts are evident in the spectrogram of LFP activity at the transient synchrony working point. (f) Joint distribution of the
frequencies and durations of these gamma bursts. Note the predominance of short-lived variable-frequency bursts.

We identified and characterized a regime of local circuit dynamics, inhibition controls the overall level of synchrony of the network.
below the onset of oscillatory synchrony, that mimics this funda- Enhancing recurrent inhibition induces a graded transition toward
mental feature and reproduces the substantial fluctuations in both higher synchrony. In Figure 1c this is shown via an increase of
frequency and power seen experimentally in local field potential the probability PI of establishing local inhibitory connections (see
(LFP) recordings12,13. Supplementary Fig. 1a,b for alternative mechanisms), as captured
The dynamics of a local circuit (or region) with transient synchrony by the network synchronization index (Online Methods). Increasing
is characterized in Figure 1. A circuit is modeled as a large network of local inhibition also leads to an overall reduction of the mean fre-
excitatory and inhibitory spiking neurons with heterogeneous local quency of the collective population oscillation (Fig. 1d) and of the
connectivity (Fig. 1a), synaptic conductances and latencies (Online firing rates of individual neurons (Supplementary Fig. 1c).
Methods; parameters summarized in Table 1). Figure 1b depicts the Proportionally increasing both the strength of recurrent inhibi-
evolution over time of the average membrane potential in the local tion and the background drive (Supplementary Figs. 2 and 3) essen-
circuit, together with the spiking patterns of individual neurons and tially maintains the population frequency (Fig. 1d) and firing rates
a histogram of the multiunit activity rate (Online Methods). In the (Supplementary Fig. 1c) but increases the synchronization in a way that
following we call the population-averaged membrane potential, a becomes smoother as the adopted parameter’s heterogeneity increases
convenient descriptor of the collective neuronal activity, an LFP-like (Supplementary Fig. 2). We define two working points (Fig. 1c,d
signal18. As indicated in Figure 1b, the amplitude of the LFP trace and Supplementary Figs. 1 and 2): one below the transition to syn-
exhibits transient epochs of increased oscillatory amplitude. These chrony and a second one at a stronger synchrony level, both giving
oscillatory events are more difficult to detect when inspecting the rise to a collective oscillation frequency close to 60 Hz. Bursts of
spiking patterns of a few single units whose firing is always stochas- transient oscillatory power are visible in the LFP spectrogram at the
tic19,20. Individual neurons can fire at every phase of the ongoing working point below the transition to strong synchrony (Fig. 1e),
oscillation, as revealed by an average circular variance of ~0.6 of the while higher synchrony levels lack this transient nature and generate
overall distribution of the phases of firing. a persistently strong power peak in the spectrogram (Supplementary
In our local circuit, as in experimental observations21, the genera- Fig. 4a,b). Statistical analysis of the duration and frequency of these
tion of oscillatory activity relies on delayed, recurrent interactions, bursts (their joint distribution is shown in Fig. 1f) shows that the
either occurring through direct delayed connections or mediated working point below the transition gives rise to gamma bursts with
by inhibitory–excitatory–inhibitory loops20. The level of recurrent frequencies broadly distributed over a range between 45 and 75 Hz

 advance online publication  nature NEUROSCIENCE


a r t ic l e s

Table 1  List of default network parameters in the transient- a b LFP spectrograms (a.u.) 103
2
synchrony regime 40 Area X

Model parameter Symbol Value 1.5


60

Frequency (Hz)
Excitatory synaptic decay time τ1E 3 ms
80
Inhibitory synaptic decay time τ1I 4 ms 1
Area Y
40
Excitatory synaptic time τ2E 1 ms
60 0.5
Inhibitory synaptic time τ2I 1 ms
Mean synaptic delay d 1.5 ms 80
Area Y Area X 1 2 3
Mean synaptic excitatory conductance g sE 5 nS Time (s)
Mean synaptic inhibitory conductance g sI 200 nS c Same-frequency d Cross-frequency
burst-timing overlap burst-timing overlap
Mean input synaptic conductance g sν 3 nS
Chance level 40
Standard deviation delay σd 0.1 ms 0.5 0.45
95% c.i.

Burst-timing correlation

Frequency Area Y (Hz)


Standard deviation synaptic excitatory conductance s 1 nS 0.4 50
g sE 0.35

s 0.3
Standard deviation synaptic inhibitory conductance 10 nS 60 0.25
g Is
0.2
© 2017 Nature America, Inc., part of Springer Nature. All rights reserved.

Standard deviation input synaptic conductance s 1 nS 0.15


g vs 70
Probability of local inhibitory connection PI 0.2 0.1
0.05
Probability of local excitatory connection PE 0.3 0 80
40 50 60 70 80 40 50 60 70 80
Probability of interareal excitatory connection PELR 0.08 Frequency (Hz) Frequency Area X (Hz)
  (two-regions model)
Probability of feedforward excitatory connection   PEff 0.08 Figure 2  Co-emergence and frequency tracking of gamma bursts.
  (three-regions model) (a) Scheme of a canonical circuit of two brain regions reciprocally
connected through long-range excitation (dashed lines). (b) Spectrograms
Probability of feedback excitatory connection PEfb 0.01
of LFP activity for coupled circuits in the transient synchrony regime. The
  (three-regions model)
times and frequencies of emergent coordinated bursts match in the two
Poisson input rate ν 10·PI kHz regions. (c) Burst-timing normalized overlap between same-frequency
gamma bursts compared with chance level (gray). Shaded band in blue
and a short mean duration close to 100 ms (i.e., four to six oscillation denotes 95% c.i. (d) Burst-timing overlaps between oscillatory bursts with
cycles). We call this the ‘transient synchrony regime’. It represents a arbitrary frequencies. Strict frequency-matching is not required for burst-
multidimensional range of parameters (cf. Supplementary Figs. 1–3) timing coordination.
leading to a qualitatively similar mixed dynamic in which asynchro-
nous and synchronous epochs both stochastically occur. Its properties relatively high values, peaking at ~0.5 for nearby collective oscilla-
are in good agreement with in vivo LFP recordings12,13. tion frequencies of 60 Hz. A similar frequency-tracking behavior has
In the following we consider model networks with transient syn- been observed experimentally22. We find that coordination between
chronies, representing multiple brain regions and linked into simple oscillatory episodes occurs not only between bursts with an exactly
structural-connectivity motifs, and analyze their emergent dynamics. matching main frequency (Fig. 2c) but also in a cross-frequency
In particular, we study their coordinated oscillatory bursting dynamics manner (Fig. 2d).
and their phase locking properties and, finally, test whether these fea- Beyond time and frequency coordination, we find that oscillatory
tures can subserve the selective and flexible routing of information. bursts also exhibit transient phase synchronization (Fig. 3). We per-
formed a time-resolved cross-covariance (XC) analysis of the LFP sig-
Simultaneous emergence of phase-locked gamma bursts nals from the two coupled regions of the structural motif in Figure 2a
We first treated the simplest case of two local circuits (corresponding (Online Methods). Transiently rising values of XC denote the onset
to two generic areas or regions X and Y), coupled by long-range exci- of epochs of increased inter-region oscillatory coherence (Fig. 3a).
tatory projections (Fig. 2a) established with equal probability (PELR ) The time dependent maximum of XC (denoted XC*(t)) occurs at a
and strength in both directions (i.e., X-to-Y and Y-to-X). If the dura- fluctuating time-lag τ*(t), generally different from zero (in-phase)
tion and the frequency of oscillatory epochs fluctuate independ- and from the average half-period (antiphase), indicating out-of-phase
ently in each brain region, it would be extremely unlikely to observe locking, as observed empirically23. The relative phase of the rhythms
simultaneous oscillatory bursts in a source and a target region aligned of the regions, ∆Φ (Online Methods), is a strongly fluctuating quan-
in a well-defined phase relation. To analyze the inter-regional tity whose distribution, ρ(∆Φ), is broad (Fig. 3b).
coordination (or lack of thereof) we examined the relative timing, The precision of phase-locking markedly increases with the instan-
frequencies and phase relationships of gamma bursts in the two taneous level of cross-covariance (Fig. 3c). As shown by the joint
connected regions. distribution of ∆Φ and XC* (Fig. 3c), events of higher-value XC* are
We found that our model spontaneously generates correlated less frequent than those with lower values, but they tend to occur at
episodes of matched timing and with tracking frequencies between specific relative phases. Conditioning the joint distribution on the
stochastically emitted gamma bursts in the two connected regions. values of XC* better reveals this phenomenon (Fig. 3c). For low val-
This is depicted in the spectrograms in Figure 2b. The burst-tracking ues, corresponding to a poorly synchronized baseline activity, the
phenomenon can be quantified by the normalized overlap between distribution of the relative phase ∆Φ remains broad (Fig. 3d). During
the occurrence times of oscillatory bursts in distinct regions as a epochs of strong cross-covariance, however, it becomes prominently
function of their instantaneous frequency (Online Methods). Figures bimodal (Fig. 3d).
2c,d reveal that, for frequencies in the gamma band, the overlap A similar preference for out-of-phase interareal phase-locking
between the timing of bursts with matching frequencies can reach is observed at the higher-synchrony working point of our system

nature NEUROSCIENCE  advance online publication 


a r t ic l e s

a Intermittent gamma
LFP cross-covariance (XC)
c log(P(∆Φ, XC*)) log(P(∆Φ|XC*)) the efficiency and direction of inter-region information exchange of
1 0 −4 −1 spontaneous activity depending on the system’s states by means of
−10
−3
information theoretical analyses.
Lag (ms)

A useful measure is provided by transfer entropy (TE) 26,27, here

XC*
0 0 0.5 −12
−5
computed based on LFPs from different regions (see Supplementary
10
−1 1 −20
−7 Fig. 7 for an equivalent analysis based on multiunit activity). TE, in a
0 1 2 3 0 0.5 1 0 0.5 1
model-free fashion, quantifies how much the knowledge of the past
Time (s) ∆Φ ∆Φ
activity of a putative source region improves the prediction of the
b d 0.25 ∆Φ *

∆Φ↑* ∆Φ↓* future activity of a target system; in this sense it constitutes a generali-

�(∆Φ) = 0.01
Low XC* zation of the Granger causality approach (exactly as mutual informa-
0.5 0 tion (MI) generalizes linear correlation). The information-theoretical
0.25 ∆Φ *
↑ ∆Φ↑* ∆Φ↓* 0 0.5 1
setting naturally allows conditioning the estimation of TE or MI on
∆Φ↓*
an arbitrary set of constraints. We designed state-selecting filters that
�(∆Φ) = 0.01

0.75 ∆Φ
0.5 0 ∆Φ↑* ∆Φ↓*
0.25 ∆Φ * condition information measures on transient dynamical features (for

example, on specific relative phase conditions to be fulfilled) and so
∆Φ↓* 0 0.5 1

�(∆Φ) = 0.01
0.75 ∆Φ High XC*0.5 0
define a state-specific directed functional connectivity (or, in short,
a routing state). As summarized in Figure 4b, transients of a sys-
© 2017 Nature America, Inc., part of Springer Nature. All rights reserved.

0.75 ∆Φ↓*
tem’s activity fulfilling specific state-filtering conditions are pooled,
0 0.5 1
∆Φ through time and trials, into separate statistical samples. Different
distributions of activity were obtained from these separate samples
Figure 3  Transient phase-locking. (a) Time-resolved cross-covariance (XC)
and quantified information-theoretically (Online Methods). Using
between simulated LFPs of the two coupled brain regions of Figure 2a.
Circuits are set at the same working point. Horizontal dotted lines mark this approach we probed information transfer specifically within
τ* = 0 ms (white) and t * = ± T / 2 (black) with T equal to the length of the state-filter-selected epochs.
average period. The positions of the highest XC maxima, displaced with Using a TE analysis in which the relative phase is constrained to a
respect to these guiding lines, reveal the occurrence of transient episodes narrow band, we study state-specific directed functional connectivity
of interareal phase-locking with out-of-phase relations. (b) Distribution between the collective neuronal activity of the regions X and Y (Fig. 4c).
of instantaneous interareal phase-differences between simulated LFPs,
For both directions of interaction, TE exhibits a broad peak cen-
in polar (left) and linearly unwrapped (right) histogram forms. Note the
two peaks at relative phases ∆Φ↓* > 0.5 and ∆Φ↑* < 0.5. (c) Left: joint
tered on a specific combination of the state delay τ and ∆Φ (Fig. 4d
probability distribution of the relative phase ∆Φ and the time-dependent and Online Methods). In both panels, the state delay, τ, that maxi-
maximum XC* of the cross-covariance XC. Right: the joint distribution mizes information transfer corresponds to a lag of ~4 ms, at which
is conditioned on the value of XC*. Note the enhanced peaks associated XC(t) peaks during coordinated bursting events (Fig. 3a). This opti-
with the high-synchrony events. (d) Polar (left) and linearly unwrapped mal state delay τ depends on the metastable phase locking (∆Φ↑* , ↓)
(right) histograms of relative phase, conditioned on XC* being lower (top) and therefore differs from the mean interareal transmission delay
or higher (bottom) than threshold levels (Online Methods). Relative phases
close to ∆Φ↓* or ∆Φ↑* dominate when XC* is high.
d of 1.5 ms (Supplementary Figs. 5 and 18 and “Discussion” section).
For longer state delays, TE does not rise significantly above chance
level (P < 0.05, bootstrap confidence interval (c.i.) comparison over the
(Supplementary Fig. 4). In this case, the simulated LFPs whole white-colored range in Supplementary Fig. 8a). Optimal func-
persistently oscillate and the two phase-locking configurations are tional coupling occurs at different relative phases for different direc-
stable attractors (Supplementary Fig. 4c), giving rise to a promi- tions of influence. This phenomenon is quantified by the functional
nently bimodal ∆Φ distribution (Supplementary Fig. 4d). The anisotropy index, ∆TE, proportional to (TEY→X − TEX→Y) (Fig. 4e
gamma-bursting dynamics can be viewed as transiently replicating and Online Methods). Two patches with equal absolute values, |∆TE|,
oscillatory modes that would be stable in the high-synchrony regime but different signs correspond to effectively unidirectional configu-
(Supplementary Fig. 5). rations with opposite directions (Fig. 4f), supported by the same
Spiking activity is only weakly modulated by oscillations on aver- fixed bidirectional structural connectivity. We refer to the states cor-
age. Nevertheless, we found that the phase-concentration of neuronal responding to these two configurations, filtered by the ∆Φ↑ (Y leads
firing transiently rises to higher values during high XC* transients X, 0 < ∆Φ < 0.5) and ∆Φ↓ (X leads Y and 0.5 < ∆Φ < 1) conditions, as
(Supplementary Fig. 6). The instantaneous circular variances of the dis- the Top and Bottom routing states, respectively. Information transfer
tribution of firing phases drops to ~0.3 during the strongest coherence proceeds from the region leading in phase to the region lagging in
oscillatory bursts and rises as high as ~0.8 during interburst periods. phase, as if long-range synaptic connections from the laggard to the
leader region were effectively nonfunctional.
Emergent routing states Based on analysis of the relative phase-dependency of TE in Figure 4d,e
The canonic two-regions motif of Figure 2a can be used as a model of we also adopted coarser state-filtering criteria, using solely the ∆Φ↑
two interconnected brain regions at two different levels in the cortical and ∆Φ↓ conditions. Figure 4g,h shows the net directed information
structural hierarchy, such as, for example, a sensory region and a pre- transfer in the Top (Fig. 4g) or Bottom routing states (Fig. 4h) com-
frontal cortical (cf. the scheme in Fig. 4a) interacting in an attention- pared to chance-level expectations (Online Methods). Information
modulated manner10 or a frontal and parietal pair of regions during transfer in the opposite laggard-to-leader direction falls below chance
different stages of motor planning24. Others have observed that the level, thus actively quenching information transfer.
direction of influence between such region pairs can change depend- Information transfer in the leader-to-laggard direction was highly
ing on task and behavioral stage10,25. We thus examined how the tran- significant despite the coarseness of the ∆Φ↓ and ∆Φ↑ conditions
sient dynamics of gamma bursts emerging in our models (Fig. 3) (P < 0.0001, bootstrap c.i. comparison). This indicates that the emer-
modulate the direction and strength of information flow. We quantify gence of information transfer anisotropy requires just a weak degree

 advance online publication  nature NEUROSCIENCE


a r t ic l e s

a b
X Y X leads Y Y leads X X leads Y Y leads X Top

Population
activity
X ...
... ...
SY TE/MI
Y
analysis Bottom
SX
1

State-filtering
condition
∆Φ↑ 0
1 State-dependent State-dependent
∆Φ↓ 0 distributions functional states

c d TEY→X (bits) e ∆TE g ∆Φ↑


i j ∆Φ↑
l 0.06 ∆Φ↑
0.25 1
1 1 0.25 0.2 TEY→X MISY↔X
Y Y TEX→Y MISX↔Y

TE (bits)

MI (bits)
3

TE (bits)
3 0.15
τ (ms)

τ (ms)
0 0.15 0.04
0.1
5 5 SY
0.05 0.05
7 7
−1 0 0.02
0 0.5 1 0 0.5 1 X Y Y X 0 5 10 0 5 10
∆Φ ∆Φ τ (ms) τ (ms)
© 2017 Nature America, Inc., part of Springer Nature. All rights reserved.

TEX→Y (bits) f h SX
k m
0.25 ∆Φ↓ ∆Φ↓ ∆Φ↓
1 0.25 0.2 0.06
TEY→X MISY↔X
3 TEX→Y MISX↔Y

TE (bits)
TE (bits)

MI (bits)
0.15
τ (ms)

0.15 X 0.04
X Bottom 0.1
5
0.05 0.05
7 Top
0 0.02
0 0.5 1 X Y Y X 95% c.i. 0 5 10 0 5 10
∆Φ τ (ms) τ (ms)
Chance level
∆Φ↑ ∆Φ↓

Figure 4  Information transfer during transient burst events. (a) Cartoon illustrating potentially interacting areas in a manner similar to X and Y; for
example, visual area V4 and frontal eye field (FEF). Inputs to these areas are modeled as temporally fluctuating signals SX and SY. (b) Scheme of the
state-filtering pipeline. LFP epochs fulfilling the state-selecting constraints (in this case probing the relative phase ∆Φ) are pooled, sampling separate
neural-activity distributions. State-dependent information-transfer direction and efficiency are evaluated from these distributions. (c) Scheme of the
model examined in panels d–h (see also Fig. 2a). (d) Maps of TE between LFPs of the X and Y regions, resolved as a function of ∆Φ and the state-delay τ.
White, TE values below chance. Information transfer is always significantly higher from the phase leading to the phase lagging region than vice versa.
(P < 0.05 at least over all the white-colored range, bootstrap c.i. comparison). (e) Functional anisotropy index, ∆TE. Colored lobes with different
signs indicate the existence of two distinct routing states, Top and Bottom, associated with effectively unidirectional information transfer in opposite
directions. (f) Graphical representation of the routing states as functional motifs (colors match the regions represented in c and e). (g,h) Information
transfer in different directions, in (g) a ∆Φ↑ configuration (Top state) or (h) a ∆Φ↓ configuration (Bottom state), for the optimal state delay. Boxes and
shaded bands show the 95% c.i.; shaded gray bands indicate chance-level values. (i) Scheme illustrating the input-driven routing model analyzed
in panels j–l. Input streams SX and SY are provided as inputs to every neuron in each area (see also Supplementary Fig. 13). (j,k) TE as a function of
the state delay τ for the model in i, in a Top state (j) and a Bottom state (k). The routing states, as measured by TE, are unidirectional and switch as a
function of the phase relation. (l,m) MI as a function of the state-delay, τ, between the input stream to one source region and the neuronal activity in
the target region, in a Top state (l) and in a Bottom state (m). The input to a source region only reaches the target region when the rhythm of the source
region leads in phase that of the target area.

of phase-locking. In fact, we find unidirectional transfer of informa- In particular, it is not obvious that streams of external information
tion during Top and Bottom transients to be robust, persisting within entering the circuits are routed in a fashion similar to the spontane-
extended ranges of local synaptic strength (Supplementary Fig. 9), ous activity analyzed so far. To examine this aspect in our models,
interareal coupling (Supplementary Figs. 9 and 11a) and synchroni- we introduce external streams of input currents, SX and SY, modeling
zation levels (Supplementary Figs. 10 and 11b), and we find that such incoming activity from surrounding areas or direct sensory stimula-
information transfer emerges for distinct mechanisms for the gen- tion (for a characterization of this state, see Supplementary Fig. 13).
eration of oscillations (Supplementary Fig. 12) or indicators of the We then studied how these input streams propagate through the inter-
population’s activity (Supplementary Fig. 7). We further restricted connected regions conditional on the current routing state (Fig. 4i
the analyses in Figure 4g,h to transients with peak XC* rising above and Online Methods). The addition of external input did not modify
(Supplementary Fig. 8c,d) or dropping below (Supplementary Fig. 8e) the system’s states as defined by TE, which remain markedly unidi-
different threshold levels (Online Methods). We found that informa- rectional (Fig. 4j,k). State-resolved information theoretical analysis
tion is transmitted primarily inside the bursts. The interareal syn- between the input streams SY (or SX) to the source region Y (or X) and
chrony thus works as a gain control that boosts information transfer the neuronal activity of the target region X (or Y) revealed that the
in one of the two directions when synchronization is high and limits information supplied to a source region propagated toward an inter-
it during asynchronous epochs. connected distant target, following the path imposed by the routing
states, defined by the state-specific TE analysis. The phase-relations
Flexible routing of information streams between the neuronal rhythms define the direction of information
The existence of a phase-gated dynamic switching between Top and flow of input signals, as revealed by the MI analysis of Figure 4l,m
Bottom routing states only starts to address the question of routing. and the TE analysis of Supplementary Figure 14.

nature NEUROSCIENCE  advance online publication 


a r t ic l e s

a Z leads X leads Y Z leads Y leads X


X leads Y leads Z

... ... ...


Population
X TE/MI
activity Y ... ... analysis

... ... ...


∆Φ2XYZ
State-filtering
condition

∆Φ3ZXY

∆Φ4ZYX State-dependent State-dependent


distributions functional states

c Y
d
∆Φ1XZY ∆Φ2XYZ ∆Φ3ZXY ∆Φ4ZYX ∆Φ5YXZ ∆Φ6YZX
b X Z
X Y
SX

Y MIS
e f MIS Z Y
© 2017 Nature America, Inc., part of Springer Nature. All rights reserved.

Z Y
MIS MIS
0.04 X Y ∆Φ2XYZ X Y
∆Φ4ZYX
X Z

MI (bits)
SZ
Z 0.02

SX SZ 0 95% c.i.
0 5 10 0 5 10
τ (ms) τ (ms) Chance level

Figure 5  Flexible routing of input signals through a hierarchy of areas. (a) Scheme showing the state-filtering pipeline for three interconnected regions,
analogous to the two-regions case described in Figure 4b. Six state filters are defined for all permutations of X, Y and Z. (b) Cartoon of a macaque brain
with three recoding sites: brown and blue in primary visual cortex V1 and red in extrastriate cortex, for example, V4. Also indicated: screen with two
stimuli (slashed circles), fixation spot (×) and the receptive fields of the recording sites in matched colors. This experimental configuration provides
a possible exemplification of the three-regions model studied here. (c) Scheme of three populations, modeling the two V1 and the V4 sites as circuits
in the transient synchrony regime. The connectivity is asymmetric, with stronger feedforward (i.e., V1 to V4) excitatory projections. (d) Functional
motifs for the associated state selecting filters. Configurations in which Y is lagging in phase behind the other areas (∆Φ1XZY and ∆Φ3ZXY) give rise to a
convergent functional motif in which both regions transmit information to Y, although with different time-courses (Supplementary Fig. 14). Functional
motifs in which Y leads both X and Z in phase are functionally disconnected, reflecting the weakness of the back-projections. Functional motifs
associated with ∆Φ2XYZ and ∆Φ4ZYX are functionally unidirectional despite the symmetrical connectivity to Y from the lower order regions X and Z.
(e) As in c except that additionally, each region in the model can receive independently fluctuating input signals (Online Methods). For clarity, only SX
and SZ are shown. (f) Left: MI between the input signal SX to region X, one of the two possible source regions, and the LFP activity of a target region Y
(as a function of the state delay; blue curve) is strongly enhanced in the associated phase configuration ∆Φ2XYZ. This phase configuration, in which X
leads Y in phase, implements a routing state mimicking attention directed toward the stimulus in the blue receptive field. No significant information
can be retrieved from Y about the signal SZ, representing the nonattended stimulus, as evidenced by the second MI curve (brown) at baseline levels.
Right: phase configuration ∆Φ4ZYX, implementing a routing state that mimics attention directed to the stimulus in the brown receptive field. MI between
the LFP activity of the target region Y and the input signal SZ is now highly significant. The routing states associated with each phase configuration are
represented by the functional motifs drawn next to the MI plots.

Next, we further extended the model to include three intercon- A routing-state analysis of the three-region system (Fig. 5c) is
nected regions, X, Y and Z. As shown in Figure 5a, the state-selection shown in Figure 5d and Supplementary Figure 15. In this setting,
procedure can be extended straightforwardly to this case. In this sys- X and Z send convergent projections toward a third higher-order
tem, six different conditions can arise, denoted by the state-selecting region, Y, which, in its turn, sends weaker feedback projections
YXZ
filters ∆Φ1XZY to ∆Φ6 (where ∆Φ1XZY indicates a first-phase con- to X and Z (for a system with equally strong back-projections see
figuration in which X leads Z leads Y within a cycle; Supplementary Supplementary Fig. 15). TE analysis of the LFP signals of the three
Fig. 15 and Online Methods). Depending on parameters and inputs, coupled regions shows that multiple routing states can arise, depending
a multitude of possible functional motifs can arise. on the phase pattern. In particular, in the routing states defined by the
This three-region architecture enabled us to study a setting typical phase configurations ∆Φ2XYZ and ∆Φ4ZYX, information transfer along
of visual selective-attention experiments11,28,29. For instance, one can one of the possible feedforward pathways is completely blocked.
consider a situation in which two distinct visual stimuli are processed This allows selection of only one of the two possible information
by distinct populations X and Z in lower-order areas such as parts of sources when external input streams are supplied to the system (Fig. 5e).
V1, whose distinct receptive fields both fall within the larger field We focus here on routing of the signals SX and SZ, coding for
of a third population, Y, in a higher-order area such as V4 (Fig. 5b). different stimuli (for more routing examples see Supplementary
In such an experiment, the focus of attention may be alternatively Fig. 16). In a phase configuration in which X leads Y leads Z, only
directed to one stimulus or the other, defining an information-rout- information about the input stream SX is present in the down-
ing problem in which the transmission toward Y of the input stream stream region Y (Fig. 5f), although no other mechanism besides an
entering X (or Z) and associated with the attended stimulus must unsuitable phase relation prevents information about SZ from arriv-
be enhanced, while the other, unattended, stream entering Z (or X) ing at Y. Switching between the routing states detected by the ∆Φ2XYZ
should be gated off. and ∆Φ4ZYX conditions thus produces rerouting effects analogous

 advance online publication  nature NEUROSCIENCE


a r t ic l e s

Y Y Y

Y X Y X Y X X Z X Z X Z

∆�Y = 4 kHz ∆�Y = 2 kHz ∆�X,Y = 0 kHz ∆�X = 2 kHz ∆�X = 4 kHz ∆�X = 4 kHz ∆�X = 2 kHz ∆�X,Z = 0 kHz ∆�Z = 2 kHz ∆�Z = 4 kHz

a d

∆ΦXY ∆ΦXY ∆ΦXY ∆ΦXY ∆ΦXY


b �(∆Φ) = 0.01
∆TE
1 1

3
τ (ms)

0
5

7
−1
0 0.5 1 0 0.5 1 0 0.5 1 0 0.5 1 0 0.5 1 ∆ΦZY ∆ΦZY ∆ΦZY ∆ΦZY ∆ΦZY
∆ΦXY ∆ΦXY ∆ΦXY ∆ΦXY ∆ΦXY

c
© 2017 Nature America, Inc., part of Springer Nature. All rights reserved.

Fraction of time in which X leads Y Fraction of time in which Z leads Y


Fraction of time Fraction of time Fraction of time in which Y leads X Fraction of time in which Y leads Z
∆Φ in which Y leads X ∆Φ in which X leads Y

Figure 6  Steering information transfer. (a) The application of a constant bias is modeled as an increase ∆ν of the background drive νin of only one of the
two interconnected regions. As a result, the phase distribution of phase relations is modified. Middle position: no input bias. Extreme positions: strong
background bias ∆ν = +4 kHz (in addition to a baseline drive of ν = 3 kHz) applied to population X (extreme right) or to Y (extreme left). Intermediate
columns: weaker cases of ∆ν = +2 kHz. (b) The predominant direction of information transfer is biased toward the region with lower background drive,
as indicated by the maps of the functional anisotropy index, ∆TE. (c) Pie charts of the time the system spends in a Top or a Bottom state. The functional
motifs for the corresponding routing states are indicated below. (d) As in a but for a three-regions model. Center: without background bias, the dynamics
of a three-region system spends the same fraction of time in a state in which both X leads Y or Z leads Y (note that there is already a weak asymmetry
in the baseline distributions of ∆ΦXY and ∆ΦZY, reflecting the different strengths of feedforward and feedback connectivity; cf. Fig. 5c). When a mild
bias is applied to population X (or Z; left and right positions respectively), the fraction of time in which X leads Y and Y leads Z (or, correspondingly,
in which Z leads Y and Y leads X) increases, eventually suppressing the fraction of time spent in the mirror configuration. Information is preferentially
transmitted to the target downstream region Y from the region receiving the background bias. See Supplementary Figure 17 for details.

to those that would be associated with a reorientation of the applied bias (extreme right or left columns) eventually mutes informa-
attentional spotlight. tion transfer in the subdominant direction.

Steering information routes DISCUSSION


In the examples of Figs. 4 and 5, transients of the different routing Our results demonstrate that, in coupled circuits with transient syn-
states (Top and Bottom for the two-regions model and ∆Φ1XZY to chrony, frequency tracking and out-of-phase locking of oscillatory
∆Φ6YXZ for the three-regions model) arise spontaneously, with simi- bursts are emergent features of the large-scale circuit’s collective
lar probabilities of occurrence. We found that applying a weak bias, dynamics. Bursts are generated in coordinated sets of near-simulta-
making the bursts of a given type more likely to occur, can enforce a neous onset and mediate transient phase-locking between the regions’
dominant direction of information transfer (Fig. 6). activities. The transient patterns of inter-regional coherence define
Figure 6a–c refers to routing-state control in the two-regions model flexible routing states that determine the direction of information
(Fig. 4), and Figure 6d refers to one in the three-regions model (Fig. 5). flow of both endogenously generated rate fluctuations and externally
Raising the baseline input to either one of the two possible source supplied signals.
regions by ∆ν increases the probability that the biased region will While the mechanisms of transient interareal frequency tracking
transiently become a phase leader, favoring the occurrence of spe- and phase-locking in biologically plausible models were largely unex-
cific routing states. As indicated by the pie charts in Figure 6c, plored, evidence for such patterns of coherence has been reported
a bias to circuit X (or Y) increases the probability of occurrence of experimentally22 (but questioned elsewhere15). Distinct from the
a Bottom (or Top) state, as seen in the rightmost (or leftmost) col- typical behavior of chains of weakly coupled oscillators 30, burst
umns. Equivalently for the three-regions model, a bias to the region coordination in our model is fast enough to conciliate transient
X increases the probability of X leading Y and, simultaneously, of phase-locking with a rapid dissolution of oscillatory synchrony after
Y leading Z (Fig. 6d). Similarly, a bias to the region Z increases the a few cycles. This intrinsic and spontaneous coordination is achieved
probability of Z leading Y and, simultaneously, of Y leading X (right despite the sparseness of interareal connections and the heterogene-
columns). As the relative phase-distributions (Fig. 6b and ity of conduction delays. Our work demonstrates that fast dynamic
Supplementary Fig. 17) take on more asymmetric shapes, maps of frequency-matching is a robust emergent feature of interacting neu-
the functional anisotropy index reveal a strong asymmetry in the ronal populations with transient synchrony and does not require an
direction of information transmission, showing that information external controller.
flows preferentially from the biased region toward the target one. Our study of the dynamics of a general regime of neuronal net-
This is detailed in Figure 6b for the two-regions model in and in work activity with transient synchrony sheds a new light on long-
Supplementary Figure 17 for the three-regions model. The highest standing arguments against coherence-mediated communication.

nature NEUROSCIENCE  advance online publication 


a r t ic l e s

A first widespread concern has been the low and inconsistent power of multifrequency interactions or multiareal networks of greater
gamma oscillations in vivo12,13. In our models, LFP oscillations show complexity. Our study lays out a methodology that can be used to
inconsistent power and fluctuate stochastically between 40 and 70 Hz, assess the dynamic routing properties of such models stringently and
and their time-averaged LFP spectrum presents broadband gamma critically and can guide the design of future experimental studies.
modulation (Supplementary Fig. 6f), consistent with other models of Features that at first sight appear to be noncompliant with informa-
transiently synchronous neuronal activity31. Despite the weak average tion routing may actually provide the brain with a particularly flexible
gamma power, TE analyses revealed that the transient power rise dur- routing mechanism.
ing oscillatory bursts is sufficient to impact on information transfer
between areas and on the propagation of externally supplied informa- Methods
tion. A second concern relates to the low level of spiking correlations. Methods, including statements of data availability and any associated
While several studies identify marked and context-dependent oscil- accession codes and references, are available in the online version of
latory synchrony32, others find only weaker spiking correlations33. the paper.
In our simulations, spiking activity always remained highly irregular.
Note: Any Supplementary Information and Source Data files are available in the
In particular, neuronal firing was poorly phase-concentrated, with
online version of the paper.
the exception of burst events (Supplementary Fig. 6). These short
epochs nevertheless convey the largest net contribution to directed Acknowledgments
information transfer (Supplementary Fig. 8c,d). A third concern is We thank U. Ernst, C.M. Gray, A. Kreiter, K. Pawelzik and R. Shapley for discussions.
© 2017 Nature America, Inc., part of Springer Nature. All rights reserved.

This work was partially supported by the Federal Ministry for Education and
raised by the large variability of interareal conduction delays, imped-
Research (BMBF) under grant no. 01GQ1005B (to A.P., T.G., F.W. and D.B.), by a
ing the phase-matched arrival of presynaptic spikes with respect to GGNB Excellence Stipend of the University of Göttingen (to A.P.), through CRC
the oscillation in the target region. The regulation of information 889 by the Deutsche Forschungsgemeinschaft and by the VolkswagenStiftung
transfer, however, is the outcome of collective emergent states and under grant no. ZN2632 (to F.W.), and by the FP7 Marie Curie career development
not just of direct monosynaptic interactions. In fact, observed inter- fellowship IEF 330792 (DynViB) (to D.B.).
areal phase relations are largely independent of the distance between AUTHOR CONTRIBUTIONS
the interacting sites34. The average interareal cross-correlation lag A.P. performed the simulations of the models and analyzed the results;
is not determined solely by the neuronal conduction delay but also A.P., F.W. and D.B. conceived the study, designed models and developed analysis
depends on the phase response properties of the coupled popula- pipelines; A.P., T.G., F.W. and D.B. wrote the paper. All authors discussed the results
and implications.
tions35,36, which in turn reflect local circuit parameters37. As shown in
Supplementary Figure 5, different interareal phase-locking relations COMPETING FINANCIAL INTERESTS
can be implemented with a fixed interareal delay just by changing the The authors declare no competing financial interests.
level of local inhibition. It is thus conceivable that the fine control of
Reprints and permissions information is available online at http://www.nature.com/
local inhibition exerted by specialized interneuron types38 might play
reprints/index.html. Publisher’s note: Springer Nature remains neutral with regard to
an active role in setting interareal phase relationships. jurisdictional claims in published maps and institutional affiliations.
Our model predicts that transitions between alternative routing
states, multiplexed within a fixed structural circuit39, are reflected by 1. Olshausen, B.A., Anderson, C.H. & Van Essen, D.C. A neurobiological model of
switching patterns of interareal phase differences40. Such a routing visual attention and invariant pattern recognition based on dynamic routing of
information. J. Neurosci. 13, 4700–4719 (1993).
mechanism would be compatible with both rate coding and more- 2. Vogels, T.P. & Abbott, L.F. Gating multiple signals through detailed balance of
complex representations relying on finer organization of the spiking excitation and inhibition in spiking networks. Nat. Neurosci. 12, 483–491
(2009).
patterns41,42. Given the irregular spiking in our models, coordinated 3. Zylberberg, A., Fernández Slezak, D., Roelfsema, P.R., Dehaene, S. & Sigman, M.
excitability fluctuations may gate the transmission of complex spike The brain’s router: a cortical network model of serial processing in the primate
patterns. Future studies could, for instance, explore whether the brain. PLoS Comput. Biol. 6, e1000765 (2010).
4. Abeles, M., Hayon, G. & Lehmann, D. Modeling compositionality by dynamic binding
simultaneous emergence of synchronized bursts can modulate infor- of synfire chains. J. Comput. Neurosci. 17, 179–201 (2004).
mation propagation through embedded systems of interacting synfire 5. Kumar, A., Rotter, S. & Aertsen, A. Conditions for propagating synchronous spiking
chains5,6 or whether stimulus-dependent cell assemblies43,44 could be and asynchronous firing rates in a cortical network model. J. Neurosci. 28,
5268–5280 (2008).
selectively gated by interareal phase coordination. 6. Hahn, G., Bujan, A.F., Frégnac, Y., Aertsen, A. & Kumar, A. Communication through
The phase relations we find may be potentially useful in establish- resonance in spiking neuronal networks. PLoS Comput. Biol. 10, e1003811–
e1003816 (2014).
ing dynamic functional hierarchies of interareal communication23, 7. Akam, T. & Kullmann, D.M. Oscillations and filtering networks support flexible
with information flowing preferentially from phase-leading toward routing of information. Neuron 67, 308–320 (2010).
phase-lagging regions. Since the structures of these phase patterns are 8. Harnack, D., Ernst, U.A. & Pawelzik, K.R. A model for attentional information
routing through coherence predicts biased competition and multistable perception.
not solely prescribed by connectivity, different regions can compete J. Neurophysiol. 114, 1593–1605 (2015).
flexibly, assuming the functional roles of sender or receiver of infor- 9. Fries, P. A mechanism for cognitive dynamics: neuronal communication through
mation, depending on the specific context and modulatory signals neuronal coherence. Trends Cogn. Sci. 9, 474–480 (2005).
10. Gregoriou, G.G., Gotts, S.J., Zhou, H. & Desimone, R. High-frequency, long-range
(Fig. 6 and Supplementary Fig. 17). coupling between prefrontal and visual cortex during attention. Science 324,
Interareal interactions may be mediated not only by coherence in 1207–1210 (2009).
11. Grothe, I., Neitzel, S.D., Mandon, S. & Kreiter, A.K. Switching neuronal inputs by
the gamma band but also by that in the beta band45 or by coordinated differential modulations of gamma-band phase-coherence. J. Neurosci. 32,
faster and slower oscillations, as in the hippocampus46,47. For simplic- 16172–16180 (2012).
ity, our model was designed to generate transient oscillations within 12. Burns, S.P., Xing, D. & Shapley, R.M. Is gamma-band activity in the local field potential
of V1 cortex a “clock” or filtered noise? J. Neurosci. 31, 9658–9664 (2011).
a single gamma band. This specification, however, is not inherent to 13. Xing, D. et al. Stochastic generation of gamma-band activity in primary visual cortex
our approach and can be relaxed to reproduce multifrequency oscilla- of awake and anesthetized monkeys. J. Neurosci. 32, 13873–13880a (2012).
tions48. At a brain-wide scale, functional networks associated with mul- 14. Jia, X., Tanabe, S. & Kohn, A. γ and the coordination of spiking activity in early
visual cortex. Neuron 77, 762–774 (2013).
tifrequency spectral fingerprints may be actively generated49,50. Future 15. Ray, S. & Maunsell, J.H.R. Do gamma oscillations play a role in cerebral cortex?
studies could thus also examine more complex models, including Trends Cogn. Sci. 19, 78–85 (2015).

 advance online publication  nature NEUROSCIENCE


a r t ic l e s

16. Ray, S. & Maunsell, J.H.R. Differences in gamma frequencies across visual cortex 34. Canolty, R.T. et al. Oscillatory phase coupling coordinates anatomically dispersed
restrict their possible use in computation. Neuron 67, 885–896 (2010). functional cell assemblies. Proc. Natl. Acad. Sci. USA 107, 17356–17361 (2010).
17. Jia, X., Xing, D. & Kohn, A. No consistent relationship between gamma power and 35. Witt, A. et al. Controlling the oscillation phase through precisely timed closed-loop
peak frequency in macaque primary visual cortex. J. Neurosci. 33, 17–25 optogenetic stimulation: a computational study. Front. Neural Circuits 7, 49
(2013). (2013).
18. Okun, M. et al. Diverse coupling of neurons to populations in sensory cortex. Nature 36. Tiesinga, P.H. & Sejnowski, T.J. Mechanisms for phase shifting in cortical networks
521, 511–515 (2015). and their role in communication through coherence. Front. Hum. Neurosci. 4, 196
19. Brunel, N. & Hakim, V. Fast global oscillations in networks of integrate-and-fire (2010).
neurons with low firing rates. Neural Comput. 11, 1621–1671 (1999). 37. Battaglia, D., Brunel, N. & Hansel, D. Temporal decorrelation of collective oscillations
20. Brunel, N. & Wang, X.-J. What determines the frequency of fast network oscillations in neural networks with local inhibition and long-range excitation. Phys. Rev. Lett.
with irregular neural discharges? I. Synaptic dynamics and excitation-inhibition 99, 238106 (2007).
balance. J. Neurophysiol. 90, 415–430 (2003). 38. Burkhalter, A. Many specialists for suppressing cortical excitation. Front. Neurosci.
21. Bartos, M., Vida, I. & Jonas, P. Synaptic mechanisms of synchronized gamma 2, 155–167 (2008).
oscillations in inhibitory interneuron networks. Nat. Rev. Neurosci. 8, 45–56 39. Battaglia, D., Witt, A., Wolf, F. & Geisel, T. Dynamic effective connectivity of inter-
(2007). areal brain circuits. PLoS Comput. Biol. 8, e1002438 (2012).
22. Roberts, M.J. et al. Robust gamma coherence between macaque V1 and V2 by 40. Dotson, N.M., Salazar, R.F. & Gray, C.M. Frontoparietal correlation dynamics reveal
dynamic frequency matching. Neuron 78, 523–536 (2013). interplay between integration and segregation during visual working memory.
23. Bastos, A.M., Vezoli, J. & Fries, P. Communication through coherence with inter- J. Neurosci. 34, 13600–13613 (2014).
areal delays. Curr. Opin. Neurobiol. 31, 173–180 (2015). 41. Osborne, L.C., Palmer, S.E., Lisberger, S.G. & Bialek, W. The neural basis for
24. Chakrabarti, S., Martinez-Vazquez, P. & Gail, A. Synchronization patterns suggest combinatorial coding in a cortical population response. J. Neurosci. 28,
different functional organization in parietal reach region and dorsal premotor cortex. 13522–13531 (2008).
J. Neurophysiol. 112, 3138–3153 (2014). 42. Rigotti, M. et al. The importance of mixed selectivity in complex cognitive tasks.
© 2017 Nature America, Inc., part of Springer Nature. All rights reserved.

25. Buschman, T.J. & Miller, E.K. Top-down versus bottom-up control of attention in Nature 497, 585–590 (2013).
the prefrontal and posterior parietal cortices. Science 315, 1860–1862 (2007). 43. Harris, K.D., Csicsvari, J., Hirase, H., Dragoi, G. & Buzsáki, G. Organization of cell
26. Schreiber, T. Measuring information transfer. Phys. Rev. Lett. 85, 461–464 assemblies in the hippocampus. Nature 424, 552–556 (2003).
(2000). 44. Buzsáki, G. Neural syntax: cell assemblies, synapsembles, and readers. Neuron 68,
27. Wibral, M., Vicente, R. & Lizier, J.T. Directed Information Measures in Neuroscience 362–385 (2010).
(Springer, 2014). 45. Salazar, R.F., Dotson, N.M., Bressler, S.L. & Gray, C.M. Content-specific fronto-
28. Bosman, C.A. et al. Attentional stimulus selection through selective synchronization parietal synchronization during visual working memory. Science 338, 1097–1100
between monkey visual areas. Neuron 75, 875–888 (2012). (2012).
29. Grothe, I. et al. Attention selectively gates afferent signal transmission to area V4. 46. Lisman, J. The theta/gamma discrete phase code occuring during the hippocampal
Preprint at https://doi.org/10.1101/019547 (2015). phase precession may be a more general brain coding scheme. Hippocampus 15,
30. Somers, D. & Kopell, N. Rapid synchronization through fast threshold modulation. 913–922 (2005).
Biol. Cybern. 68, 393–407 (1993). 47. Colgin, L.L. et al. Frequency of gamma oscillations routes flow of information in
31. Mazzoni, A., Panzeri, S., Logothetis, N.K. & Brunel, N. Encoding of naturalistic the hippocampus. Nature 462, 353–357 (2009).
stimuli by local field potential spectra in networks of excitatory and inhibitory 48. Cannon, J. et al. Neurosystems: brain rhythms and cognitive processing. Eur. J.
neurons. PLoS Comput. Biol. 4, e1000239 (2008). Neurosci. 39, 705–719 (2014).
32. Kreiter, A.K. & Singer, W. Stimulus-dependent synchronization of neuronal responses 49. Hipp, J.F., Hawellek, D.J., Corbetta, M., Siegel, M. & Engel, A.K. Large-scale cortical
in the visual cortex of the awake macaque monkey. J. Neurosci. 16, 2381–2396 correlation structure of spontaneous oscillatory activity. Nat. Neurosci. 15, 884–890
(1996). (2012).
33. Ecker, A.S. et al. Decorrelated neuronal firing in cortical microcircuits. Science 50. Bastos, A.M. et al. Visual areas exert feedforward and feedback influences through
327, 584–587 (2010). distinct frequency channels. Neuron 85, 390–401 (2015).

nature NEUROSCIENCE  advance online publication 


ONLINE METHODS Here
N
Network models. The networks used to model local brain regions consisted of ∑ i = 1Vi (t )
1,000 excitatory and 250 inhibitory conductance-based Wang-Buzsáki neurons51 V (t ) =
with random connectivity. Synaptic conductances had a difference-of-exponen- N
tials time-course, proportional to a postsynaptic channel open probability Ps: is the LFP-like signal52,53. The variables s V (t ) and s Vi (t ) denote the s.d. of LFP
amplitudes over time or, respectively, of the membrane potential traces Vi(t) of
 − t − t0 − d −
t − t0 − d 
each individual neuron i. The χ2 coefficient is bounded to the unit interval, with
 t1 t2  (1)
Ps = B e −e  (1) vanishing values indicating asynchronous dynamics. The frequency of the col-
lective oscillation of a region is calculated as the inverse of the first peak position
where t0 is the time of the presynaptic spike, d is a combined axonal and synaptic of the autocorrelation function of a multiunit rate signal, obtained by convolving
transmission delay, τ1 and τ2 define the rise and decay synaptic time-constants the raster plot of each local network with a Gaussian kernel of s.d. of 2 ms. The
(τrise = τ1τ2/(τ1 − τ2) and τfall = τ1), respectively, and B is a normalization fac- rate signal in Figure 1b was calculated as a histogram with a 1-ms bin width.
tor. Excitatory (E) or inhibitory (I) synaptic currents were then given by Single neuron rates were calculated as the inverse of their mean interspike inter-
Is(E,I) = gs(E,I)Ps(E,I)(Vs(E,I) − V), where gs(E,I) is the peak conductance of the con- val. Spectrograms of LFP activity (Figs. 1e and 2c and Supplementary Figs. 4a
sidered synapse and Vs(E,I) is the synaptic reversal potential. We used VsE = 0 mV and 13b) were calculated using a standard time–frequency Fourier analysis with
and VsI = −80 mV for excitatory and inhibitory synapses, respectively. overlapping Hamming windows (300-ms size, 250-ms overlap) and the time-
Both peak synaptic conductances and synaptic delays were Gaussian-distrib- averaged spectra through the multitaper method (16 tapers).
uted random variables with prescribed means gs(E,I) and d and s.d. sg ( E , I ) and
© 2017 Nature America, Inc., part of Springer Nature. All rights reserved.

s
σd. The probability of establishing an excitatory (or inhibitory) connection within Analyses of gamma-burst properties and coordination. To extract the joint
each modeled brain region is denoted by PE (or PI). The background drive to each distribution of frequencies and durations of oscillatory burst (Fig. 1f and
neuron in the network was provided by external excitatory inputs, modeled as Supplementary Fig. 4b), we thresholded LFP spectrograms (300-ms size, 280-ms
Poisson processes with rate νin and statistically independent for different neurons. overlap) at the 95th percentile of instantaneous power values. The thresholded
Each spike injected by this external driving source induced in the target neuron a spectrograms were then scanned at different frequencies (i.e., line by line) to
synaptic current as in equation (1). Peak conductances of the external drive were extract the durations of time intervals over which power at the given frequency
also heterogeneous and assumed values sampled from a truncated Gaussian dis- was continuously sustained above threshold. We sampled these burst durations
n
tribution with mean g s and s.d. sg n . Long range synaptic connectivity between by combining 20 trials (overall, 100 min of signal). To study time and frequency
s
two or three brain regions was purely excitatory, targeting both excitatory and coordination between oscillatory bursts occurring in different coupled regions
inhibitory neurons in the target region. In the two-regions network architec- (Fig. 2c–d), LFP spectrograms of each region were thresholded above their 95th
ture, the probability of having a long-range excitatory projection, PELR, was held percentile and scanned line by line, creating a binary string for each popula-
equal between regions. In the three-regions network architecture, the region tion and for each frequency z X ,Y ( f ) . These binary strings were then multiplied
Y (at the highest hierarchical level) received feedforward connections from X against the analogous strings obtained from the other population and normalized
and Z with a probability PE ff that was stronger than or equal to probability PE fb by the s.d. of the binary strings to obtain the overlap
of feedback connections from Y to both X and Z. Both feedforward (and z ( f ) ⋅z X ( f )
feedback) connections had equal conductance strength. No connection was J( f ) = Y
s z X s zY
present between X and Z. Long-range excitatory synaptic conductance values
were drawn from the same distribution as for recurrent excitation within each The calculation was made using 50 trials for Figure 2c and 10 trials for
local region. Supplementary Figures 4 and 13. Significance of the estimated overlap (Fig. 2c)
The default set of default network parameters is summarized in Table 1. These was assessed by comparing 95% c.i. for each overlap value with the associated
default values were used, unless specified otherwise (specific parameter variations permutation-based chance levels (2,500 replicas).
are listed in the following). Simulations were performed with PI = 0.3 in Figures 1,
2 and 6 and Supplementary Figure 17; PI = 0.6 in Supplementary Figure 4; and Analysis of transient phase-locking. We calculated time-resolved cross-cov-
PI = 0.1 in Supplementary Figures 12 and 14f–h. The probability of establish- ariance (XC) adopting an overlapping sliding window 50 ms long (roughly two
ing a local excitatory connection deviated from default values in Figures 4 and 5 average oscillation periods) and with 0.1-ms steps. Within each time-window,
and Supplementary Figures 15–17 with PE = 0.35; in Supplementary Figures 7 standard cross-covariance was performed between chunks of LFPs of each of
and 14 with PE = 0.25; and in Supplementary Figure 12 with PE = 0. In Figure 6, the coupled populations:
we also adopted a smaller degree of input conductance heterogeneity, sg n = 0.1
s
nS; in Supplementary Figures 9, 10, 13 and 16 we adopted sg n = 0.5 nS; and XC(t , d ) = 〈(VX (t ′) − VX ) ⋅ (VY (t ′ − d ) − VY )〉t ′ ∈W (t ) (2) (2)
s
in Supplementary Figures 15 and 17 we adopted sg n = 0 nS. Large values of
s
σg ν might lead to symmetry-breaking network instances, in which one specific
s
where 〈⋅〉t ′∈W (t ) denotes averaging over a 50-ms time-window W(t) centered on
phase-locking configuration is favored over of the other even when there is no bias time t, and VX and VY are LFP averages over this same window.
applied. For clarity of illustration, these symmetry-breaking instances were excluded The instantaneous phase of each LFP signal ΦX and ΦY was calculated by inter-
from all analyses. Typical simulation runs were performed with a fourth-order polating a straight line over the unit phase interval 0 < Φ < 1 between consecutive
Runge-Kutta method when the input was a smooth random field (fixed time-step maxima of the population activity of a single repetition. A peak in population
∆t = 0.05 ms or simple Poisson spike-trains (fixed time-step ∆t = 0.1 ms), and lasted activity was considered such if it was the maximum in a neighborhood of a half-
10 min unless otherwise indicated (details found in the corresponding section). period radius of the central oscillation frequency of the population activity. Such
When the input was modeled as an Ornstein-Uhlenbeck process, an Euler- a fast approach avoids bandpass-filtering and leads to similar results as a Hilbert
Maruyama integration scheme with ∆t = 0.01 ms was used instead. To sample transform analysis, as previously shown35. The relative phase is then defined by
neuronal activity distributions, we combined multiple simulations with quenched ∆Φ = (ΦY − ΦX, mod 1). The joint probability distribution of the time-dependent
heterogeneity and connectivity and different realizations of the stochastic drive to phase difference ∆Φ and the time dependent peak of XC, XC*(t) = maxδ(t, δ), was
the system. Generally, no statistical methods were used to predetermine sample sizes, obtained from 100 simulated trials (overall, 1,000 min of signal). The conditional
but our sample sizes are similar to those reported in previous publications27,39. probability distributions shown in Figure 3C were obtained by independently
normalizing each line of the probability matrix in Figure 3c. The XC* threshold
Characterization of network activity. The synchronization index was chosen for Figure 3d was the 99th percentile (XC* = 0.37).
to be the previously introduced χ2, where The circular variance of Supplementary Figure 6 was calculated as follows.
NsV2 (t ) Given {sk}X with k = 1, …, K, the set of spike-times within a window of time
c2 = N W(t) centered at t, emitted by neurons in subset X of K neurons belonging to
∑ sV2 (t )
i =1 i the region X, we define θkX(t) as the phases of the LFP of area X at spike-times

nature NEUROSCIENCE doi:10.1038/nn.4569


θkX(t) = ΦX(sk). The local circular variance of the firing of the subset X, CVX(t), time instants in which specific conditions hold. TE and pTE are then evaluated
is then the circular variance of the set of phases θkX(t). We then average circular in a consistent regime, specified by the choice of the filtering conditions, Ξ.
variances over the two coupled regions, i.e., Information-theoretical analysis with a specific filtering condition based on the
instantaneous interareal phase-difference ∆Φ returns the routing state.
1
CV (t ) =
2
(
CVX (t ) + CVY (t ) ) When considering the two-regions model in Figures 4d,e and 6 and
Supplementary Figures 8a–c and 17, the ∆Φ range is split into five relative-
We evaluated these circular variances considering two different conditions. phase bins of equal size (the TE vs. τ and the ∆Φ surfaces were then smoothed
In Supplementary Figure 5c–e, the subsets X and Y included all the neurons by linear interpolation). Figure 6 and Supplementary Figure 17 show the
belonging, respectively, to the regions X and Y. In the right panels of anisotropy index from single trials. The filtering criteria in just two Top and
Supplementary Figure 5d,e, these sets were then restricted to just ten ran- Bottom routing states (Fig. 4g–m and Supplementary Figures 7, 8d,e, 9–12,
domly chosen neurons within the regions. In both cases, the length of the sliding 14 and 17) were selected time instants in which the relative phase fell, respectively, in
windows W(t) was equal to that of seven average oscillation periods. We used the coarser ranges ∆Φ↑, i.e., 0 < ∆Φ < 0.5, for the Top state or ∆Φ↓, i.e., 0.5 < ∆Φ < 1,
200 noise realizations with the same network architecture, lasting 1 min each, for the Bottom state.
to obtain the joint distribution of CV(t) and XC*(t). As before, the conditional In Supplementary Figure 8, we adopted a state conditional to the strength
probability distributions shown in Supplementary Figure 6d,e were obtained by of interareal cross-covariance, in addition to the instantaneous phase relation.
normalizing each line of the probability matrix in Supplementary Figure 6d,e. Conditions on the instantaneous value of XC* were imposed arbitrarily, requiring
it to exceed a threshold XC*high = median(XC*) + 0.5 (Supplementary Fig. 8b,d)
Analyses of information transfer. We evaluated TE26,27 between discrete time- or to be less than XC*low = median(XC*) − 0.5 (Supplementary Fig. 8b,e).
© 2017 Nature America, Inc., part of Springer Nature. All rights reserved.

series of the simulated LFP signals VX(t) and VY(t). These signals are constrained In the three-regions model in which a pairwise analysis was devised (Fig. 6
to take values from the set Ξ, and have distributions p(xt) and p(yt). The set Ξ and Supplementary Fig. 17), the procedure was identical to that described above
corresponds to the set of signal values observed within specific temporal windows for the two-regions model. More complex state-filtering conditions were devised
in which certain conditions, defining a consistent functional state, are properly for complete analysis of the dynamics of the three regions. The six possible hierar-
fulfilled (state-selecting filter; see below). TE from region X to region Y is defined chical configurations that the LFP signals can take define six filtering conditions,
as a deviation from the Markov condition54,55: p( yt | yt(k ) , xt(l ) ) = p( yt | yt(k ) ). ∆Φ1XZY to ∆Φ6YZX , giving rise to different (or sometimes the same) routing
This reads: the probability of VY taking a value yt at time t, given that the k past states, as illustrated in Figure 4d. The conditions can be written in terms of the
values of VY were yt(k ) = ( yt − t1 , …, yt − t k ) , is independent from the l past pairwise phase relations:
values of VX given by xt(l ). This condition is only fulfilled if there is no influence
def
from the past values of VX on the current values of VY ; in other words, only ∆Φ1XZY = ∆Φ XY > 0.5, ∆Φ XZ > 0.5, ∆Φ XZ > ∆Φ XY
if the distributions are conditionally independent. These probabilities are not
def
time dependent; they depend only on the time lags (τ1, …, τk). TE can then be ∆Φ2XYZ = ∆Φ XY > 0.5, ∆Φ XZ > 0.5, ∆Φ XZ < ∆Φ XY
expressed as a measure of the divergence from the above Markov condition: def
 p( yt | y(k ) , x(l ) ) 
∆Φ3ZXY = ∆ΦYZ < 0.5, ∆Φ XZ < 0.5, ∆Φ XZ < ∆ΦYZ (6)
TE X →Y (t1 , …, t k ) = ∑ p( yt , yt(k ) , xt(l ) )log 2  t t
 (3) def
yt ,…, yt − t k  p( y | y(k ) ) 
t t ∆Φ 4ZYX = ∆ΦYZ < 0.5, ∆Φ XZ < 0.5, ∆Φ XZ > ∆ΦYZ
∈Ξ
xt ,…, xt − t l def
(3) ∆ΦYXZ
5 = ∆Φ XY < 0.5, ∆Φ ZY < 0.5, ∆Φ XY < ∆Φ ZY
def
Here we relaxed the Markov condition and adopt a commonly used simpli­ ∆ΦYZX
6 = ∆Φ XY > 0.5, ∆Φ ZY > 0.5, ∆Φ XY > ∆Φ ZY (6)
fication39,56–62. We replaced the past activity history yt(k ) and xt(l ) in the previous
expression with unique observations of past values at a single latency at For the two-regions model, we estimated histograms (within the two possible
time yt(k ) = yt − t and xt(k ) = xt − t , and then treated the interaction latency τ Top and Bottom routing states) in Figure 4g,h and Supplementary Figure 8d–e
as a variable parameter. With this approximation, the TE expression in based on 500 min of simulated signal from 50 trials of the same network instance;
equation (3) can be simplified as in Figures 4d,e and 6b and Supplementary Figures 8c and 17b based on
10 min of signal from a single network realization (the mean in Supplementary
TE X →Y (t ) = H  yt | yt − t  − H  yt | xt − t , yt − t  (4) (4) Fig. 8a is based on a total of 150 min from 15 different network instances); for
   
Supplementary Figures 9 and 10 based on single trial simulations, each lasting
TE represents the reduction in uncertainty (entropy, H) about the future state 10 min; in Supplementary Figure 11 based on 40 different network instances
of a given region Y when the history of region X is also taken into account. An (10 min of LFP signal for each considered parameter configuration); in
analogous (but not identical) expression for TE in the opposite Y-to-X direction Supplementary Figure 12 based on 10-min simulations from 50 different network
is obtained by exchanging the indices X and Y. In Supplementary Figure 18, instances (in Supplementary Fig. 7, only 10 network instances were used, with the
we compared the results obtained with the simplified expression of equation same simulation time); and in Figure 4j–m and Supplementary Figure 14 based
(4) against estimations of TE considering different state-delays for sender and on simulations lasting 50 min each from 40 different network instances (2,000 min
receiver or considering more than a single state-delay. of simulation time). In the three-regions model, we estimated histograms (within
When more than two regions constitute the interacting system, the above each of the six state-filtering conditions) in Figure 5d and Supplementary Figure 15
expression for TE must be modified to account for the possibility of indirect based on simulations lasting 3,600 min each from 12 different network instances;
functional interactions (for example, transfer from X to Z via Y, etc.). We used in Figure 5f and Supplementary Figure 16 based on 20 network instances last-
partial TE63 to uncover the direct interactions between pairs of populations when ing 4,000 min in total; and in Figure 6 and Supplementary Figure 17 based on
embedded in a larger multiareal system. Given a set of three populations X, Y and simulations lasting 10 min each from 25 different network realizations.
Z, partial TE (pTE) with a single state-delay is defined as TE values obtained from the sampled histograms using the definition in equa-
tion (4) were then further corrected against finite-size bias by quadratic extrapola-
pTE X →Y (t ) = H[ yt | yt − t , zt − t ] − H[ yt | xt − t , yt − t , zt − t ] (5) (5)
tion toward infinite sample size64. Given a set of time series of length L, we took
This expression reduces to equation (4) when the activity of Z is statistically three subsets of the same length Lk, with Lk = L/k with k = 1, … , 10. For each
independent from that of X and Y. trimmed length Lk, TE is calculated from each of the three subsets and then aver-
All the probability densities required for the evaluation of equations (4) and aged over them to obtain a value TEk. Finally, we evaluated the limit to infinite
(5) can be straightforwardly derived from the three-term or four-term joint prob- sample size for TE, extrapolating it as the vertical-axis intercept of a second-order
ability distribution p( yt , xt − t , yt − t ) in the TE case or p( yt , xt − t , yt − t , zt − t ) polynomial fit of TEk to k. The 95% c.i. for TE values were approximated as ±2× s.d.
in the partial TE case. These probabilities must be sampled uniquely over the above and below the mean, over the considered network instances or trials (95%

doi:10.1038/nn.4569 nature NEUROSCIENCE


rule). These c.i. were then compared to the null hypothesis (lack of functional signal SX to the source region X and LFP from a target region Y can be state-
coupling), providing a baseline TE range for significance testing. To obtain the filtered as TE by pooling only input- and response-value pairs within the epochs
null-hypothesis distribution of TE, we relied on resampling techniques and used fulfilling the filtering criteria specified by the chosen filter, Ξ, between the
bootstrapping with replacement to build a distribution of 1,500 surrogates for LFP signals VX and VY:
network realization, and then collected them. To maintain the oscillatory nature
of the analyzed time-series, we built bootstrap surrogates using LFP blocks of a  p( yt , st − t ) 
mean length . This value should be long enough to guarantee that the signal’s
MISX ↔Y (t ) = ∑ p( yt , st − t )log 2  
st , yt ∈ Ξ  p( yt ) p(st − t ) 
autocorrelation structure would be preserved when building bootstrap surro-
gates. To choose , we analyzed the behavior of the bootstrapped TE versus the When analyzing three-region structural motifs in Figure 5 and Supplementary
chunk length, . For oscillatory signals in the transient synchrony regime, the Figure 16, for TE, we computed partial (also called conditional) MI65,66, intro-
value of the bootstrapped TE initially increases and then saturates at around  = ducing an additional condition on the third region signal.
100 ms. We thus chose  = 300 ms for the TE baseline calculations, and we kept it
the same for the MI calculations, although the random inputs are not oscillatory Statistics. Generally, we used nonparametric statistical testing procedures based
and therefore the values of the bootstrapped MI were independent of . After on comparisons between confidence intervals obtained by bootstrapping with
building a null-hypothesis histogram of TE values, we determined the range replacement (for both the association and the null hypotheses). See details in
between its 2.5th and 97.5th percentiles and used it as the baseline TE range. each specific Online Methods subsection. A Supplementary Methods Checklist
TE values whose 95% c.i. was completely disjoint and fell above (or below) this is available.
baseline range were indicative of significantly enhanced (or damped) information
© 2017 Nature America, Inc., part of Springer Nature. All rights reserved.

transfer with respect to chance level. Code availability. Custom-written code for model simulations and their analyses
Finally, we defined the functional anisotropy index: are available on reasonable request to the corresponding authors.

TEY → X (t ) − TE X → Y (t ) Data availability. All simulations analyzed during the study can be regenerated
∆TE(t ) =
max[| TE X →Y (t ) |,| TEY → X (t ) |] using the above code.

bounded in the −1 ≤ ∆TE ≤ 1 range, as a simpler indicator of the different effi-


ciencies of information transfer in the two possible reciprocal communication
51. Wang, X.J. & Buzsáki, G. Gamma oscillation by synaptic inhibition in a hippocampal
directions between a pair of regions. interneuronal network model. J. Neurosci. 16, 6402–6413 (1996).
52. Golomb, D. & Rinzel, J. Clustering in globally coupled inhibitory neurons. Physica
Modeling of the external input signals. Exogenous sources of information D 72, 259–282 (1994).
were modeled with two different input models. In the first approach (Fig. 5 and 53. Golomb, D. & Hansel, D. The number of synaptic inputs and the synchrony of large,
sparse neuronal networks. Neural Comput. 12, 1095–1139 (2000).
Supplementary Figs. 14 and 16), external inputs were modeled as scalar Gaussian 54. Vicente, R., Wibral, M., Lindner, M. & Pipa, G. Transfer entropy--a model-free
random field (GRF) input currents. GRFs were generated as follows. First, we measure of effective connectivity for the neurosciences. J. Comput. Neurosci. 30,
built a scalar random field in Fourier representation, given by 45–67 (2011).
55. Wibral, M. et al. Measuring information-transfer delays. PLoS One 8, e55809
p ˆ (2013).
yˆ (k) = C(k)( Ak + i Bk ) (7) (7)
∆k 56. Gourévitch, B. & Eggermont, J.J. Evaluating information transfer between auditory
cortical neurons. J. Neurophysiol. 97, 2533–2543 (2007).
where Ak and Bk are independent Gaussian random variables with 0 mean and 57. Honey, C.J., Kötter, R., Breakspear, M. & Sporns, O. Network structure of cerebral
unit variance. The correlation function Cˆ (k) in Fourier space was obtained by the cortex shapes functional connectivity on multiple time scales. Proc. Natl. Acad.
Sci. USA 104, 10240–10245 (2007).
discrete Fourier transformation of a set of T points sampled from a continuously 58. Lungarella, M., Pitti, A. & Kuniyoshi, Y. Information transfer at multiple scales.
smooth time-domain representation, C(τ) = 2/cosh(τ/τs). The T samples of C(τ) Phys. Rev. E 76, 056117 (2007).
were taken at regularly spaced times tn = n∆τ, n = 1, … , T, resulting in samples of 59. Garofalo, M., Nieus, T., Massobrio, P. & Martinoia, S. Evaluation of the performance
Cˆ (k) at k = k1, … , kT, evenly spaced with ∆k = 2π/∆τT. Finally, the real part of the of information theory-based methods and cross-correlation to estimate the functional
connectivity in cortical networks. PLoS One 4, e6482 (2009).
inverse Fourier transform of equation (7) provided the desired GRF variable. 60. Ito, S. et al. Extending transfer entropy improves identification of effective
The currents conveying the externally supplied inputs S(t) = gΨ(t) were gener- connectivity in a spiking cortical network model. PLoS One 6, e27431
ated independently for each area. Instances of S(t) were first generated in Matlab (2011).
and then fed as input currents to each neuron within the corresponding region. 61. Stetter, O., Battaglia, D., Soriano, J. & Geisel, T. Model-free reconstruction of
excitatory neuronal connectivity from calcium imaging signals. PLoS Comput. Biol.
We used the same fourth-order Runge-Kutta integrator (∆t = 0.05 ms) as described 8, e1002653 (2012).
above. The parameters used for the simulations of Figure 5 and Supplementary 62. Orlandi, J.G., Stetter, O., Soriano, J., Geisel, T. & Battaglia, D. Transfer entropy
Figures 14 and 16 were τs = 1 ms, with a gain factor of g = 2. Qualitatively similar reconstruction and labeling of neuronal connections from simulated calcium
results were obtained over the entire range 0.5 ms ≤ τs ≤ 5 ms. imaging. PLoS One 9, e98842 (2014).
63. Vakorin, V.A., Krakovska, O.A. & McIntosh, A.R. Confounding effects of indirect
In Supplementary Figure 14e–h, we used a different noise model, modeling connections on causality estimation. J. Neurosci. Methods 184, 152–160
inputs as Ornstein-Uhlenbeck processes. The input signal was given by dS(t)/dt = (2009).
−γS(t) + gη(t), where η is a Gaussian white noise with 0 mean and unit variance. 64. Strong, S., Koberle, R., de Ruyter van Steveninck, R. & Bialek, W. Entropy and
In this figure, we used g = 0.1 and γ = 1. An Euler-Mayurama method was used to information in neural spike trains. Phys. Rev. Lett. 80, 197 (1998).
65. Cover, T.M. & Thomas, J.A. Elements of Information Theory 2nd edn. (Wiley-
simulate the input and the network dynamics, with a time step ∆t = 0.01 ms. Interscience, 2006).
Information sharing between the two endpoints of a communication 66. Frenzel, S. & Pompe, B. Partial mutual information for coupling analysis of
line was assessed via the classical measure MI65. MI between an injected multivariate time series. Phys. Rev. Lett. 99, 204101 (2007).

nature NEUROSCIENCE doi:10.1038/nn.4569

You might also like