s41598-021-91244-w

Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

www.nature.

com/scientificreports

OPEN A convolutional neural network


for estimating synaptic
connectivity from spike trains
Daisuke Endo1,9, Ryota Kobayashi2,3,4,9, Ramon Bartolo5, Bruno B. Averbeck5,
Yasuko Sugase‑Miyamoto6, Kazuko Hayashi6,7, Kenji Kawano6, Barry J. Richmond5 &
Shigeru Shinomoto1,8*
The recent increase in reliable, simultaneous high channel count extracellular recordings is exciting
for physiologists and theoreticians because it offers the possibility of reconstructing the underlying
neuronal circuits. We recently presented a method of inferring this circuit connectivity from neuronal
spike trains by applying the generalized linear model to cross-correlograms. Although the algorithm
can do a good job of circuit reconstruction, the parameters need to be carefully tuned for each
individual dataset. Here we present another method using a Convolutional Neural Network for
Estimating synaptic Connectivity from spike trains. After adaptation to huge amounts of simulated
data, this method robustly captures the specific feature of monosynaptic impact in a noisy cross-
correlogram. There are no user-adjustable parameters. With this new method, we have constructed
diagrams of neuronal circuits recorded in several cortical areas of monkeys.

More than half a century ago, Perkel, Gerstein, and ­Moore1 pointed out that by measuring the influence of one
neuron on another through a cross-correlogram, physiologists could infer the strength of the connection between
the neurons. If this were done for lots of pairs of neurons, a map of the neuronal circuitry could be built. Now,
with the advent of high-quality simultaneous recording from large arrays of neurons, it might have become pos-
sible to map the structures of neuronal circuits.
The original cross-correlation method can give plausible inferences about connections. However, in many
cases, it also tended to suggest the presence of connections that are spurious, i.e., false positives (FPs). There
were many possible sources for the lack of reliability and specificity, such as large fluctuations produced by
external signals or higher-order interactions among neurons. Over the years, there have been many attempts to
minimize the presence of such spurious connections, by shuffling spike t­ rains2, by jittering spike t­ imes3–6, or by
taking fluctuating inputs into a­ ccount7–13. These, in general, helped eliminate the FPs, but they then tended to
be conservative, giving rise to false negatives (FNs), i.e., missing existing connections.
Recently, we developed an estimation method by applying the generalized linear model (GLM) to each
cross-correlogram14. The estimation method we call GLMCC works well in balancing the conflicting demands
of reducing FPs and reducing FNs, demonstrating that the cross-correlogram image actually contains sufficient
information from which to infer the presence of monosynaptic connectivity. GLMCC nonetheless has a short-
coming: the estimation results are sensitive to the model parameters, and therefore the parameters need to be
tuned for the spiking data.
Here, we develop another method: Convolutional Neural Network for Estimating synaptic Connectivity
from spike Trains (CoNNECT). The premise is that a convolutional neural network is good at capturing the fea-
tures important for distinguishing among different categories of i­ mages15–18; we apply it to a cross-correlogram,
expecting that it is capable of detecting the signature of the monosynaptic impact in the one-dimensional cross-
correlogram image. Our new method CoNNECT is easy to use, and it works robustly with data arising from
different cortical regions in non-human primates. The convolutional neural network has tens of thousands of

1
Graduate School of Informatics, Kyoto University, Kyoto 606‑8501, Japan. 2Mathematics and Informatics
Center, The University of Tokyo, Tokyo 113‑8656, Japan. 3Department of Complexity Science and Engineering,
The University of Tokyo, Chiba 277‑8561, Japan. 4JST, PRESTO, Saitama 332‑0012, Japan. 5Laboratory of
Neuropsychology, NIMH/NIH/DHHS, Bethesda, MD 20814, USA. 6Human Informatics and Interaction Research
Institute, National Institute of Advanced Industrial Science and Technology, Tsukuba 305‑8568, Japan. 7Japan
Society for the Promotion of Science, Tokyo 102‑0083, Japan. 8Brain Information Communication Research
Laboratory Group, ATR Institute International, Kyoto 619‑0288, Japan. 9These authors contributed equally:
Daisuke Endo and Ryota Kobayashi. *email: [email protected]

Scientific Reports | (2021) 11:12087 | https://doi.org/10.1038/s41598-021-91244-w 1

Vol.:(0123456789)
www.nature.com/scientificreports/

internal parameters. The parameters can be adjusted using hundreds of thousands of pairs of spike trains gener-
ated with a large-scale simulation of the circuitry of realistic model neurons. To reproduce large fluctuations in
real spike trains, we added external fluctuations to the model neurons in the simulation.
CoNNECT promptly provides reasonable inference. It does not, however, give a rationale for why the result
was derived, whereas our previous algorithm GLMCC does because it fits an interaction kernel to the cross-
correlogram. These methods, therefore, have different strengths and weaknesses and can be used in combination
in a complementary manner. Namely, the inference given by CoNNECT can be used for guiding GLMCC to
search for suitable parameters, and GLMCC can provide interpretation.
We evaluated the accuracy of estimation by comparing the inference with the true connections, using syn-
thetic data generated by simulating circuitries of model neurons, and compared the performance of CoNNECT
with that of GLMCC, as well as the classical cross-correlogram m ­ ethod19,20, the Jittering m
­ ethod4,5, and an
extended GLM ­method13. After confirming the performance of the model, we applied CoNNECT to parallel spike
signals recorded from three cortical areas of monkeys and obtained estimation of the local neuronal circuitry
among many neurons. We have found that the connections among recorded units are sparse; they are less than
1% for all three datasets.

Results
Training and validating with synthetic data. CoNNECT infers the presence or absence of monosyn-
aptic connections between a pair of neurons and estimates the amplitude of the postsynaptic potential (PSP) that
one neuron would drive in another. The estimation is performed by applying a convolutional neural ­network15–18
to a cross-correlogram obtained for every pair of spike trains (Fig. 1a). The network has an output layer con-
sisting of two units. One unit indicates the presence or absence of connectivity with a real value z ∈ [0, 1] by
thresholding at 0.5. Another is the level of PSP represented in a unit of (mV). The network was trained with
spike trains generated by a numerical simulation of a network of multiple-timescale adaptive threshold (MAT)
model ­neurons21–25 interacting through fixed synapses. In a large-scale simulation, we applied fluctuating inputs
to a subset of neurons to reproduce large fluctuations in real spike trains in vivo (Fig. 1b). Figure 1c,d, and e
demonstrate sample spike trains, histograms of the firing rates of excitatory and inhibitory ­neurons26, and firing
irregularity measured in terms of the local variation of the interspike intervals Lv27,28. The training data does not
contain many low firing rate neurons, considering the situation that low firing units are often discarded when
analyzing real data. The details of the learning procedure are summarized in “Methods” section.
We validated the estimation performance of CoNNECT using novel spike trains generated by another neu-
ronal circuit with different connections. Figure 2a depicts an estimated connection matrix, referenced to the
true connection matrix, of 50 neurons. Here, the estimation was done with spike trains recorded for 120 min. Of
50 spike trains, 40 and 10 are, respectively, sampled from 800 excitatory and 200 inhibitory neurons. Figure 2b
compares the estimated PSPs against true values. We have presented an estimated PSP as being 0 if the connection
is not detected. Points lying on the nonzero x-axis are existing connections that were not detected or FNs. Points
lying on the nonzero y-axis are spurious connections assigned for unconnected pairs or FPs. Figure 2c depicts
how the numbers of FNs and FPs for excitatory and inhibitory categories changed with the recording duration
or the length of spike trains (10, 30, and 120 min). While the number of FPs or spurious connections does not
depend largely on the recording duration, the number of FNs or missing connections decreased with the period,
implying that more synaptic connections of weaker strength are revealed by increasing the recording time.

Comparison with other estimation methods. There are many algorithms that were developed to estimate syn-
aptic connections from spike ­trains1–13,19,29–31. We compared CoNNECT with the conventional cross-correlation
method (CC)19, Jittering ­method4, Extended ­GLM13, and ­GLMCC21 for their ability to estimate connectivity
using synthetic data. Figure 3 shows connection matrices determined by the four methods referenced to the
true connection matrices. In the lower panels, we demonstrated the performances in terms of the false positive
(discovery) rate (FPR) and false negative (omission) rate (FNR) for excitatory and inhibitory categories; smaller
values are better. Here we estimated the mean and SD of the performance by applying each method to 8 test data-
sets of 50 neurons. Overall performance with FPR and FNR was measured in terms of the Matthews correlation
coefficient (MCC) (see “Methods” section). The MCCs for these estimation methods are shown in the right edge
panel; a larger MCC is better. For evaluating the performances, we adopted spiking data generated by a network
of MAT models and a network of Hodgkin–Huxley (HH) type models (“Methods” section). In computing the
numbers of FPs and FNs, we ignored small excitatory connections, which are inherently difficult to discern with
this observation duration. We took the lower thresholds as 0.1 mV for the MAT simulation and 1 mV for the HH
simulation so that the visible connectivity is about 10%, but the relative performances between different models
are unchanged even if we change the thresholds.
The conventional cross-correlation analysis produced many FPs, revealing a vulnerability to fluctuations in
cross-correlograms. The Jittering method succeeded in avoiding FPs but missed many existing connections,
thus generating many FNs. The Extended GLM method of given parameters was also rather conservative. In
comparison to these methods, GLMCC and CoNNECT have better performance, producing a small number of
FPs and FNs and a larger MCC value. Here we have modified GLMCC so that it achieves higher performance
than the original ­algorithm14 by using the likelihood ratio test to determine the statistical significance (“Methods”
section). When comparing these two algorithms, GLMCC was slightly conservative, producing more FNs, while
CoNNECT tended to suggest more connections, producing more FPs.
The converted GLMCC was better than CoNNECT for the HH model data (Fig. 3b), but the converse was
true for the MAT model data (Fig. 3a). This might be because CoNNECT was trained using the MAT model data
of a similar kind, and GLMCC was constructed by considering the HH model simulation. Although the model

Scientific Reports | (2021) 11:12087 | https://doi.org/10.1038/s41598-021-91244-w 2

Vol:.(1234567890)
www.nature.com/scientificreports/

CoNNECT
Cross-Correlogram
presence/absence
of connectivity

t post-synaptic
potential (PSP)

1D Convolution Fully connected Output


Pooling
+ tanh activation layer layer

b d
300

external fluctuations
excitatory
200 neurons
7 Hz

number
10 Hz
100
inhibitory
20 Hz neurons

0
0 5 10 15 20
firing rate [Hz]
c spike trains
e 140
120
100 inhibitory excitatory
neurons neurons
80

number
60
40
20
0
0.8 0.9 1.0 1.1 1.2 1.3
firing irregularity (Lv)

Figure 1.  The architecture of CoNNECT. (a) The algorithm infers the presence or absence of monosynaptic
connectivity and the value of postsynaptic potential (PSP) from the cross-correlogram obtained from a pair of
spike trains. The figure of a monkey was illustrated by Kai Shinomoto and licenced to Springer Nature Limited.
(b) The algorithm is trained with spike trains generated by a numerical simulation of neurons interacting
through fixed synapses. Slow fluctuations were added to a subset of neurons to reproduce large fluctuations in
real spike trains in vivo. (c) Sample spike trains (cyan: inhibitory neurons; magenta: excitatory neurons). (d)
Firing rates of excitatory and inhibitory neurons. (e) Firing irregularity measured in terms of the local variation
of the interspike intervals Lv.

performance was examined with independent datasets, the HH model simulations would be more similar than
across the HH model and MAT model. As the GLMCC parameters were selected with the HH model simulation,
it naturally works better for the HH data than MAT simulation data and vice versa.

Comparison of different learning conditions. While the convolutional neural network has the advantage that
tens of thousands of parameters can be suitably adjusted to reproduce given datasets, it does not guarantee the
generalization capability. We evaluated the generalization capability of our convolutional network by changing
the number of out-channels representing the degree of system adaptability from 1 to 10. Figure 4a depicts the
numbers of FPs and FNs in the above, and the overall performances measured in terms of MCC in the below,
which were obtained for the MAT model simulation data (left panel) and the Hodgkin-Huxley model simula-
tion data (right panel). We observe that the convolutional network consisting of 1 channel slightly “under-fits”
because of the little flexibility, whereas that of 10 channels slightly over-fits the data, exhibiting slightly lower
MCC. Thus we have employed the network of 5 channels, consisting of about fifty thousand parameters.

Scientific Reports | (2021) 11:12087 | https://doi.org/10.1038/s41598-021-91244-w 3

Vol.:(0123456789)
www.nature.com/scientificreports/

a Estimated connectivity True connectivity


1 Presynaptic neuron index N 1 Presynaptic neuron index N
1 1

Postsynaptic neuron index

Postsynaptic neuron index


N N

b Original and estimated PSPs c FPs and FNs


2 120

FNs (exc)
100 FNs (inh)
FPs (exc)
Estimated PSP [mV]

1 80 FPs (inh)

60

0 40

20

-1 0
-1 0 1 2 10 min 30 min 120 min
Original PSP [mV] Recording time

Figure 2.  Synaptic connections estimated using CoNNECT. (a) An estimated connection matrix, referenced to
a true connection matrix. Of 50 neurons, 40 and 10 are excitatory and inhibitory neurons sampled from 1000
model neurons simulated for 120 min. Excitatory and inhibitory connections are represented by magenta and
cyan squares of the sizes proportional to the postsynaptic potential (PSP). (b) Estimated PSPs plotted against
true parameters. Points on the nonzero y-axis represent the false positives (FPs) for unconnected pairs. Points
on the nonzero x-axis represent the false negatives (FNs). (c) The numbers of FPs and FNs for excitatory and
inhibitory categories counted for different recording durations.

Selection of training data sets. To make the convolutional network applicable to data of a wider variety, we
have trained the network using the cross-correlograms augmented by rescaling the time. Figure 4b depicts the
estimation performances of networks trained using the cross-correlations rescaled by 1/4 and 1/2 times of the
original (indicated as “1/4” and “1/2”), the original cross-correlations (“1”), and all the data (“1/4 + 1/2 + 1”). The
networks trained with lower firing rates exhibited lower performances. We have adopted the network trained
with all the data (“1/4 + 1/2 + 1”) because it gave the highest performance in estimating connectivity.

Cross‑correlograms. To observe the situations in which different estimation methods succeeded or failed in
detecting the presence or absence of synaptic connectivity, we examined sample cross-correlograms of neuron
pairs of a network of MAT model neurons. Figure 5 depicts neuron pairs that exhibited various patterns includ-
ing pathological cases. The majority of neuron pairs are of successful cases demonstrated in the upper part of
the figure. Some cross-correlograms from this simulation exhibited large fluctuations that resemble what is seen

Scientific Reports | (2021) 11:12087 | https://doi.org/10.1038/s41598-021-91244-w 4

Vol:.(1234567890)
www.nature.com/scientificreports/

a MAT model simulation


Cross-Correlation Jittering ExGLM GLMCC CoNNECT True connectivity

FPR + FNR: 0.28 FPR + FNR: 0.72 FPR + FNR: 0.49 FPR + FNR: 0.32 FPR + FNR: 0.25
1.0 1.0 1.0 1.0 1.0 1.0

0.8 0.8 0.8 0.8 0.8 0.8

0.6 0.6 0.6 0.6 0.6 0.6

CoNNECT
0.4 0.4 0.4 0.4 0.4 0.4

GLMCC
Jittering
ExGLM
0.2 0.2 0.2 0.2 0.2 0.2

CC
0.0 0.0 0.0 0.0 0.0 0.0
FP FN FP FN FP FN FP FN FP FN FP FN FP FN FP FN FP FN FP FN
exc inh exc inh exc inh exc inh exc inh MCC

b Hodgkin-Huxley model simulation


Cross-Correlation Jittering ExGLM GLMCC CoNNECT True connectivity

FPR + FNR: 0.44 FPR + FNR: 0.95 FPR + FNR: 0.93 FPR + FNR: 0.25 FPR + FNR: 0.40
1.0 1.0 1.0 1.0 1.0 1.0

0.8 0.8 0.8 0.8 0.8 0.8

ExGLM
0.6 0.6 0.6 0.6 0.6 0.6

Jittering

CoNNECT
0.4 0.4 0.4 0.4 0.4 0.4

GLMCC
0.2 0.2 0.2 0.2 0.2 0.2

CC
0.0 0.0 0.0 0.0 0.0 0.0
FP FN FP FN FP FN FP FN FP FN FP FN FP FN FP FN FP FN FP FN
exc inh exc inh exc inh exc inh exc inh MCC

Figure 3.  Comparison of estimation methods using two kinds of synthetic data. (a) The multiple-timescale
adaptive threshold (MAT) model simulation. (b) The Hodgkin–Huxley (HH) type model simulation. estimated
using the conventional cross-correlation method (CC), Jittering method, Extended GLM (ExGLM), GLMCC,
and CoNNECT are depicted, referenced to the true connectivity of the synthetic data. Estimated connections
are depicted in equal size for the first two methods because they do not estimate the amplitude of PSP. (lower
panels) The false-positive rate (FPR) and false-negative rate (FNR) for excitatory and inhibitory categories;
smaller values are better. The mean and SD were obtained by applying each method to 8 test datasets of 50
neurons. The sum of FPR and FNR averaged over excitatory and inhibitory categories is presented above each
panel. (lower rightmost panel) Overall performances of the estimation methods compared in terms of the
Matthews correlation coefficient (MCC); the larger, the better.

in real biological data. These were produced by external fluctuations added to a subset of neurons, making the
connectivity inference difficult. The inference results obtained by the four estimation methods are distinguished
with colors; magenta, cyan, and gray represent that estimated connections were excitatory, inhibitory, or uncon-
nected, respectively. We also superimposed a GLM function fitted to each cross-correlogram.
Figure 5a,b depict sample cross-correlograms of neuron pairs that are connected with excitatory and inhibi-
tory synapses, respectively. For the first three cross-correlograms from the top, all four estimation methods
succeeded in detecting excitatory or inhibitory connections, thus making true positive (TP) estimations. For
the fourth case, the Jittering method failed to detect the connection. This implies that the Jittering method is
rather conservative for producing FPs, and as a result, has produced many FNs. In this case, the cross-correlation
method (CC) has mistaken the excitatory synapse as inhibition due to the large wavy fluctuation in the cross-
correlogram. For the last cases, all four estimation methods failed to detect the connection, resulting in FNs.
This would have been because the original connections were not strong enough to produce significant impacts
on the cross-correlograms.

Scientific Reports | (2021) 11:12087 | https://doi.org/10.1038/s41598-021-91244-w 5

Vol.:(0123456789)
www.nature.com/scientificreports/

a MAT model simulation Hodgkin-Huxley model simulation


100 100
80 FPs (exc) FNs (exc) 80
FPs (inh) FNs (inh)
60 60
40 40
20 20
0 0

1.0 1.0
0.8 0.8
0.6 0.6
MCC MCC
0.4 0.4
0.2 0.2
0.0 0.0
1 3 5 10 1 3 5 10

Number of channels Number of channels

b MAT model simulation Hodgkin-Huxley model simulation


100 100
80 FPs (exc) FNs (exc) 80
FPs (inh) FNs (inh)
60 60
40 40
20 20
0 0

1.0 1.0
0.8 0.8
0.6 0.6
MCC MCC
0.4 0.4
0.2 0.2
0.0 0.0
1/4 1/2 1 1/4+1/2+1 1/4 1/2 1 1/4+1/2+1

Firing rates of training data Firing rates of training data

Figure 4.  Comparison of different learning conditions. (a) The convolutional networks of different numbers
of channels. We have adopted a network of 5 channels. (b) The convolutional networks trained using the
cross-correlations rescaled by 1/4 and 1/2 times the original (indicated as “1/4” and “1/2”), the original cross-
correlations (“1”), and all the data (“1/4 + 1/2 + 1”). We have adopted the network trained with all the data. The
numbers of FPs and FNs for the excitatory and inhibitory categories estimating the connectivity of 50 neurons
are depicted above, and the Matthews correlation coefficient (MCC) is depicted below. The performances are
tested with the MAT model simulation data (left panel) and the Hodgkin-Huxley model simulation data (right
panel).

Figure 5c depicts sample cross-correlograms of unconnected pairs. For the first two cross-correlograms, all
four estimation methods judge the absence of connections correctly (or the null hypothesis of the absence of
connection was not rejected), resulting in true negatives (TNs). For the third pair, the CC suggested the presence
of a connection, resulting in an FP. This demonstrates that the conventional cross-correlation method is fragile
in the presence of large fluctuations. For the fourth and the last cases, the CC, GLMCC, and CoNNECT have
suggested monosynaptic connections. The sharp peaks appearing in the cross-correlogram would have been
caused by indirect interaction via other neurons. In such cases, however, it is difficult to discern the absence of
a monosynaptic connection solely from the cross-correlogram.

Analyzing experimental data. We examined spike trains recorded from the prefrontal (PF), inferior
temporal (IT), and the primary visual (V1) cortices of monkeys using the Utah arrays. Experimental conditions
of individual data are summarized in “Methods” section. Because neurons with low firing rates do not provide

Scientific Reports | (2021) 11:12087 | https://doi.org/10.1038/s41598-021-91244-w 6

Vol:.(1234567890)
www.nature.com/scientificreports/

a b c
true: excitatory true: inhibitory true: unconnected

CC Jit GLM CoNN CC Jit GLM CoNN CC Jit GLM CoNN


1200 250 120
1000 200 100
800 80
150
600 60
100
400 40
200 Pre: 19, Post: 12 50 Pre: 46, Post: 6 20 Pre: 2, Post: 21
0 0 0
CC Jit GLM CoNN CC Jit GLM CoNN CC Jit GLM CoNN
400 1000 60
800 50
300
40
600
200 30
400
20
100 200
Pre: 2, Post: 11 Pre: 50, Post: 4 10 Pre: 16, Post: 28
0 0 0
CC Jit GLM CoNN CC Jit GLM CoNN CC Jit GLM CoNN
120 500 300
100 400 250
80 200
300
60 150
200
40 100
20 100 50
Pre: 22, Post: 24 Pre: 43, Post: 11 Pre: 4, Post: 19
0 0 0

CC Jit GLM CoNN CC Jit GLM CoNN CC Jit GLM CoNN


1500 200 200
160 160
1000
120 120
80 80
500
40 40
Pre: 8, Post: 49 Pre: 44, Post: 38 Pre: 13, Post: 46
0 0 0

CC Jit GLM CoNN CC Jit GLM CoNN CC Jit GLM CoNN


250 600 120
200 500 100
400 80
150
300 60
100
200 40
50 100 Pre: 45, Post: 50 20 Pre: 23, Post: 18
Pre: 2, Post: 5
0 0 0
-50 0 50 -50 0 50 -50 0 50
s [ms] s [ms] s [ms]

Figure 5.  Sample cross-correlograms obtained from the MAT model simulation. (a), (b), and (c) Pairs of
neurons that have excitatory and inhibitory connections, and are unconnected, respectively. Four kinds of
estimation methods, the cross-correlation (CC), the Jittering (Jit), GLMCC (GLM), and CoNNECT (CoNN),
were applied to cross-correlograms. Their estimation (excitatory, inhibitory, and unconnected) are respectively
distinguished with colors (magenta, cyan, and gray). The lines plotted on the cross-correlograms are the GLM
functions fitted by GLMCC. The causal impact from a pre-neuron to a post-neuron appears on the right half in
each cross-correlogram.

enough evidence for the connectivity, we have excluded low firing units and examined those that have fired more
than 1 Hz.

Preprocessing experimental data. Some of the experimentally available cross-correlograms exhibit a sharp
drop near the origin for a few ms due to the shadowing effect, in which near-synchronous spikes cannot be
­detected32. This effect disrupts the estimation of synaptic impacts that should appear near the origin of the cross-
correlogram. The data were obtained with a sorting algorithm specifically used for the Utah array exhibit rather
broad shadowing effects larger than 1 ms (up to 1.75 ms). Here, we analyzed the experimental data by removing
an interval of 0 ± 2 ms in the cross-correlogram and applying the estimation method to a cross-correlogram
obtained by concatenating the remaining left and right parts (Fig. 6a,b). We also conducted this operation in the
analysis of synthetic data.
Figure 6c demonstrates the cross-correlograms of sample neuron pairs for which both CoNNECT and
GLMCC estimated connections which were excitatory, inhibitory, or absent (unconnected). It was observed
that the real cross-correlograms are accompanied by large fluctuations. Nevertheless, CoNNECT and GLMCC
are able to detect the likely presence or absence of synaptic interaction by ignoring the severe fluctuations.

Scientific Reports | (2021) 11:12087 | https://doi.org/10.1038/s41598-021-91244-w 7

Vol.:(0123456789)
www.nature.com/scientificreports/

c CoNNECT: + CoNNECT: - CoNNECT: 0


GLMCC: + GLMCC: - GLMCC: 0
300 180 1600
250 160 1400
140 1200
200 120 1000
150 100 800
80
100 60 600
40 400
50 Pre: 169, Post: 177 Pre: 23, Post: 54 Pre: 22, Post: 4
a 0
20
0
200
0
PF 450 300 300
400 250 250
350
300 200 200
250 150 150
200
150 100 100
100 Pre: 35, Post: 36 50 Pre: 71, Post: 72 50 Pre: 26, Post: 7
50
0 0 0
-50 0 50 -50 0 50 -50 0 50
s [ms] s [ms] s [ms]
450 800 80
400 700 70
350 600 60
300 500 50
250 400 40
200
150 300 30
100 200 20
Pre: 21, Post: 40 100 Pre: 27, Post: 28 10 Pre: 25, Post: 17
50
0 0 0
IT
b 80
70
250 600
500
60 200
50 400
150
40 300
30 100
200
20 50 Pre: 167, Post: 168
10 Pre: 42, Post: 25 100 Pre: 83, Post: 38
0 0 0
-50 0 50 -50 0 50 -50 0 50
s [ms] s [ms] s [ms]
160 450 450
140 400 400
120 350 350
100 300 300
80 250 250
200 200
60 150 150
40 100 100
20 Pre: 6, Post: 5 Pre: 34, Post: 35 Pre: 18, Post: 5
50 50
0 0 0
V1
600 300 160
-50 0 50 -50 0 50 500 250 140
s [ms] s [ms] 120
400 200 100
300 150 80
200 100 60
40
100 Pre: 47, Post: 46 50 Pre: 22, Post: 79 20 Pre: 25, Post: 22
0 0 0
-50 0 50 -50 0 50 -50 0 50
s [ms] s [ms] s [ms]

Figure 6.  Cross-correlograms of real spike trains recorded from PF, IT, and V1 using the Utah arrays. (a) An
interval of 0 ± 2 ms in the original cross-correlogram was removed to mitigate the shadowing effect, in which
near-synchronous spikes were not detected. (b) Processing real cross-correlograms. (c) The cross-correlograms
for which CoNNECT and GLMCC gave the same inference. The fitted GLM functions are superimposed on
the histograms. The causal impact from a pre-neuron to a post-neuron appears on the right half in the cross-
correlogram.

Connection matrices. Figure 7 depicts the estimated connections for the entire three datasets of PF, IT, and V1.
The units in the connection matrices are arranged in the order provided by a sorting algorithm, and accordingly,
units of neighboring indexes of the matrices tended to have been spatially closely located. All three connection
matrices had more components in near diagonal elements, implying that neurons in a nearby location are more
likely to be connected. The firing rate and irregularity (the local variation of the interspike intervals Lv27,28) are
shown in the rightmost panels. The summary statistics in Table 1 reflect differences in firing rate between excita-
tory and inhibitory cells in PF and IT but not V1. The firing irregularity of excitatory neurons is slightly higher
than that of inhibitory neurons, consistent with the previous results.
Table 1 summarizes the statistics of the three datasets. Each neuron is assigned as putative excitatory, puta-
tive inhibitory, or undetermined, according to whether the excitatory–inhibitory (E–I) dominance index is
positive, negative, or undetermined (or zero), respectively. Here, the E–I dominance index is defined as

Scientific Reports | (2021) 11:12087 | https://doi.org/10.1038/s41598-021-91244-w 8

Vol:.(1234567890)
www.nature.com/scientificreports/

PF T=120 min, N = 214, Pairs: 45582, Connected: 0.4%

Putative excitatory
Putative inhibitory
Undetermined

Firing irregularity (Lv)


1

0
0 2 4 6 8 10 12 14 16

Firing rate (Hz)

IT T = 120 min, N = 170, Pairs: 28730, Connected: 0.4%

Putative excitatory
Putative inhibitory
Undetermined

Firing irregularity (Lv)


1

0
0 2 4 6 8 10 12 14 16

Firing rate (Hz)

V1 T = 30min, N = 88, Pairs: 7656, Connected: 0.6%

2
Putative excitatory
Putative inhibitory
Undetermined
Firing irregularity (Lv)

0
0 2 4 6 8 10 12 14 16

Firing rate (Hz)

Figure 7.  Connection matrices and diagrams estimated for spike trains recorded from the prefrontal (PF),
inferior temporal (IT), and the primary visual (V1) cortices of monkeys. In the connection diagrams, excitatory
and inhibitory dominant units are depicted as triangles and circles, respectively, and units with no outgoing
connections or those that innervate equal numbers of estimated excitatory and inhibitory connections are
depicted as squares. (rightmost panels) The firing rate and irregularity (Lv ) of the putative excitatory and
inhibitory units identified by the E–I dominance index. Units that had no innervating connections or those that
exhibited vanishing E–I dominance indices are depicted in gray.

dei = (ne − ni )/(ne + ni ), in which ne and ni represent the numbers of excitatory and inhibitory identified con-
nections projecting from each unit, ­respectively14. The row “num. connections” indicates the average number of
innervated connections per neuron. Because the number of innervated connections for each neuron is only a

Scientific Reports | (2021) 11:12087 | https://doi.org/10.1038/s41598-021-91244-w 9

Vol.:(0123456789)
www.nature.com/scientificreports/

Area PF IT V1
Recording time 120 min 120 min 30 min
Neurons 214 170 88
(Putative excitatory) 95 55 28
Firing rates ± SD (Hz) 3.55 ± 2.83 3.57 ± 3.48 7.48 ± 5.82
Irregularity ± SD (Lv) 0.98 ± 0.20 1.02 ± 0.18 1.30 ± 0.21
(Putative inhibitory) 27 22 5
Firing rates ± SD (Hz) 6.12 ± 3.03 7.96 ± 5.90 3.93 ± 1.59
Irregularity ± SD (Lv) 0.85 ± 0.16 0.99 ± 0.16 1.21 ± 0.16
Num. connections 0.91 0.76 0.53
Directed pairs 45582 28730 7656
(Putative excitatory) 143 92 37
(Putative inhibitory) 52 37 10
Connecting % 0.42 0.45 0.61

Table 1.  Results of analyzing experimental datasets.

a PF IT V1

0 - 60 min 60 -120 min 0 - 60 min 60 -120 min 0 - 15 min 15 - 30 min

b
0.5 0.5 1.5
second half PSP

second half PSP

second half PSP

1.0
0.25 0.25
0.5
0.0 0.0
0.0

-0.25 -0.25
-0.5
-0.25 0.0 0.25 0.5 -0.25 0.0 0.25 0.5 -0.5 0.0 0.5 1.0 1.5
first half PSP first half PSP first half PSP

Figure 8.  Stability of connection estimation. (a) Connection matrices estimated for the first and second halves
of the spike trains recorded from PF, IT, and V1. (b) Comparison of the PSPs estimated from the first and
second halves.

few, the majority of dei is either 0 or 1. Though we have obtained many connections, the total number of all pairs
is enormous, scaling with the square of the number of units, and accordingly, the connectivity is sparse (less
than 1% for each (directed) pair of neurons).
In contrast to synthetic data, the currently available experimental data do not contain information regarding
the true connectivity. To examine the stability of the estimation, we split the recordings in half and compared
estimated connections from each half. If the real connectivity is stable, we may expect the estimated connections
have overlap between the first and second halves. Figure 8a represents the connection matrices obtained from
the first and second halves of the spike trains recorded from PF, IT, and V1. Figure 8b compares the estimated
PSPs in two periods. Many estimated connections appear only on one of the two. This might be simply due to
statistical fluctuation or due to real changes in synaptic connectivity. Nevertheless, it may be noteworthy that
the excitatory connections of large amplitudes were detected relatively consistently between the first and second
halves. Namely, they appear in the first and third quadrants diagonally, implying that they have the same signs
with similar amplitudes.

Scientific Reports | (2021) 11:12087 | https://doi.org/10.1038/s41598-021-91244-w 10

Vol:.(1234567890)
www.nature.com/scientificreports/

Discussion
Here we have devised a new method for estimating synaptic connections based on a convolutional neural net-
work. While this method does not require adjusting parameters for individual data, it robustly provides a rea-
sonable estimate of synaptic connections by tolerating large fluctuations in the data. This high performance was
obtained by training a convolutional neural network using a considerable amount of training data generated by
simulating a network of model spiking neurons subject to fluctuating current.
We compared CoNNECT with the conventional cross-correlation method, the Jittering method, Extended
GLM, and GLMCC in their ability to estimate connectivity, using synthetic data obtained by simulating neuronal
circuitries of fixed synaptic connections. Both CoNNECT and GLMCC exhibited high performance in predicting
individual synaptic connections, superior to other methods.
Then we applied CoNNECT to simultaneously recorded spike trains recorded from monkeys using the Utah
arrays. We have found that the connections among recorded units are sparse; they are less than 1% for all three
datasets. To test the reliability of the estimation, we divided the entire recording interval in half and estimated
connections for respective intervals. We have seen that strong excitatory connections overlap between the peri-
ods. This result implies that the estimation is reliable for the strong connectivity, and the connectivity lasts at
least for hours.
The cross-correlograms of real biological data (Fig. 6) turned out to be even more complicated than those
of synthetic data (Fig. 5), which were generated by adding large fluctuations to individual neurons (Fig. 1). The
complicated features in real cross-correlograms were not only due to fluctuations in real circuitry but also due
to the sorting algorithm. The most severe bottleneck in estimating connectivity may have been the shadowing
effect of a few ms, in which near-synchronous spikes were not detected (Fig. 6a); this effect might hide the first
part of a monosynaptic impact, which is expected to show up in a few ms in a cross-correlogram. If the sorting
algorithm is improved such that the shadowing duration is shortened, the estimation might be more reliable.
In this study, we have employed the convolutional neural network to capture the specific signature of mono-
synaptic impact in a cross-correlogram image. While the convolutional network is known to be robust against the
translation of images, the monosynaptic impact is expected to appear at a specific location in the cross-correlo-
gram, particularly exhibiting the delay of a few milliseconds. Thus it might be an interesting challenge to search
for other learning algorithms that utilize such information and perform better than the convolutional network.
We used data augmentation t­ echnique33 to increase the number of training examples artificially. Data aug-
mentation is known to improve the performance on various tasks in computer v­ ision18,34 and acoustic signal
­processing35,36. Here we augmented the cross-correlation data by rescaling the time to capture the diverse synaptic
interactions. This augmentation also improved the estimation performance of synaptic connectivity (Fig. 4b).
Recently, several authors proposed a more systematic approach for data augmentation, e.g., generating augmented
data using generative adversarial networks (GANs)37,38 and learning the data augmentation p ­ olicy39. Though these
approaches focus on the image classification task and require a vast computational resource, it may be interesting
to apply these techniques to pursue an advanced data augmentation method for synaptic connectivity estimation.
So far, we have little knowledge about neuronal circuitry in the brain. By collecting more data from high chan-
nel count recordings and applying these reliable analysis methods to them, we shall be able to obtain information
about neuronal circuitry in different brain regions and learn about network characteristics and the information
flow in each area. Ultimately, we expect that we will characterize the network characteristics of different brain
regions processing various kinds of information.

Methods
Configuration of a neural network for estimating synaptic connectivity. Here we describe the
details of a four-layered convolutional neural ­network15–18 applied to cross-correlograms obtained for every pair
of spike trains to estimate the presence or absence of a connection and its postsynaptic potential (PSP) (Fig. 1).
The neural network learns to find a bump or dent in the cross-correlogram caused by a monosynaptic connec-
tion.
In particular, the input consists of 100 integer values of the spike counts in a cross-correlation histogram in
an interval of [−50, 50] ms with 1 ms bin size. The network comprises a 1-dimensional convolution layer, the
average pooling, and the internal layer of fully connected 100 nodes. The output layer consists of two units; one
indicates the presence or absence of connectivity with a real value z ∈ [0, 1]. Another is the level of PSP repre-
sented in a unit of (mV).

Training the convolutional neural network. We ran a numerical simulation of a network of 1000 neurons inter-
acting through fixed synapses in various conditions and trained the neural network with spike trains from 400
units selected from the entire network. Thus, we constructed cross-correlograms of about 80, 000 pairs, each
assigned with the teaching signals consisting of the true information about the presence or absence of connec-
tivity (respectively represented as z = 1 or 0) and its PSP value in either direction. The training was performed
using an algorithm for first-order gradient-based optimization of stochastic objective functions, based on adap-
tive estimates of lower-order moments, named ­Adam40. The parameters adopted in the learning are summarized
in Table 2. Details of the architecture are summarized in Table 3. Figure 9 demonstrates a set of convolutional
kernels that were learned with training data. From the set of learned kernels, we can see some specific features
of monosynaptic impact of a few milliseconds appearing in a cross-correlogram. It is also interesting to see a
kernel exhibiting a roughly monotonic gradient (the second panel from the left). This might have worked for
detrending the large slow fluctuations in the cross correlogram produced by our simulation, which aimed at
reproducing real situations.

Scientific Reports | (2021) 11:12087 | https://doi.org/10.1038/s41598-021-91244-w 11

Vol.:(0123456789)
www.nature.com/scientificreports/

Learning rate, β1, β2 0.001, 0.9, 0.999


Dropout No
Epochs 20
Loss PSP: mean squared error
Connectivity: binary cross entropy
Loss weight PSP: 0.5, connectivity: 0.5

Table 2.  Hyperparameters of the convolutional neural network.

Convolution layer
Kernel size 10
Number of out-channels 5
Slide range 1
Activation function tanh
Hidden layer
Number of nodes 100
Activation function ReLU

Table 3.  Architecture of the convolutional neural network.

0.4 0.6 0.4 0.5


0.4
0.3 0.5 0.3 0.4
0.3
0.2 0.4 0.2 0.3
0.2
0.3 0.2
0.1 0.1 0.1
0.2 0.1
0 0 0
0.1 0
-0.1 -0.1 -0.1 -0.1
0
-0.2 -0.2 -0.2
-0.2 -0.1
-0.2 -0.3 -0.3
-0.3 -0.3
1 2 3 4 5 6 7 8 9 10 1 2 3 4 5 6 7 8 9 10 1 2 3 4 5 6 7 8 9 10 1 2 3 4 5 6 7 8 9 10 1 2 3 4 5 6 7 8 9 10
time [ms] time [ms] time [ms] time [ms] time [ms]

Figure 9.  A set of convolutional kernels that were learned from training data. The range of the kernel is 10 ms.

Data augmentation. To make the estimation method applicable to data of a wider range, we performed data
­augmentation18,34–37 ­(see33 for review). Namely, we augmented the data by rescaling the cross-correlations by 2
and 4 times and used all the data, including the original data in the learning.

Web‑application program. A ready-to-use version of the web application, the source code, and example data
sets are available at our website, https://s-​shino​moto.​com/​CONNE​CT/ and are also hosted publicly on Github,
accessible via https://​github.​com/​shige​rushi​nomoto. The simulation code is also available at this Github.

Improvement of GLMCC. Original framework of GLMCC. In the previous s­ tudy14, we developed a meth-
od of estimating the connectivity by fitting the generalized linear model to a cross-correlogram, GLMCC. We
designed the GLM function as

(1)
 
cc (t) = exp a(t) + J12 f (t) + J21 f (−t) ,

where t is the time from the spikes of the reference neuron. a(t) represents large-scale fluctuations in the cross-
correlogram in a window [−W, W] (W = 50 ms). By discretizing the time in units of �(= 1ms), a(t) is repre-
sented as a vector a� = (a1 , a2 , . . . , aM ) ( M = 2W/�). J12 ( J21) represents a possible synaptic connection from
the reference (target) neuron to the target (reference) neuron. The temporal profile of the synaptic interaction
is modeled as f (t) = exp(− t−d τ ) for t > d and f (t) = 0 otherwise, where τ is the typical timescale of synaptic
impact and d is the transmission delay. Here we have chosen τ = 4 ms, and let the synaptic delay d be selected
from 1, 2, 3, and 4 ms for each pair.
The parameters θ� = {J12 , J21 , a�} are determined with the maximum a posteriori (MAP) estimate, that is, by
maximizing the posterior distribution or its logarithm:

log p(θ�|{ti }) = log p({ti }|θ�) + log p(θ�) + const., (2)


where {ti } are the relative spike times. The log-likelihood is obtained as

Scientific Reports | (2021) 11:12087 | https://doi.org/10.1038/s41598-021-91244-w 12

Vol:.(1234567890)
www.nature.com/scientificreports/

npre nsp    W
(j) (j)
 
log p({ti }|θ�) = log cc ti − cc (t)dt, (3)
j=1 i=1 −W

where npre is the number of spikes of presynaptic neuron (j). Here we have provided the prior distribution of a
that penalizes a large gradient of a(t) and uniform prior for {J12 , J21 }
M−1
 1
log p(θ�) = − (ak+1 − ak )2 + const, (4)
γ�
k=1

where the hyperparameter γ representing the degree of flatness of a(t) was chosen as γ = 2 × 10−4 [ms−1].

Likelihood ratio test. The likely presence of the connectivity can be determined by disproving the null hypoth-
esis that a connection is absent. In the original model, this was performed by thresholding the estimated param-
eters with |Jˆij | > 1.57zα (τ Tpre post )−1/2, where zα, T, pre, and post are a threshold for the normal distribution,
recording time, firing rates of pre- and post-synaptic neurons. But we realized that this thresholding method
might induce a large asymmetry in detectability between excitatory and inhibitory connections.
Instead of a simple thresholding, here we introduce the likelihood ratio test that is a general method for testing
hypothesis (Chapter 11 ­of41, see ­also42): We compute the likelihood ratio between the presence of the connectivity
Jij = Ĵij and the absence of connectivity Jij = 0 or its logarithm:
 
D = log L∗ Jij = Ĵij − log L∗ Jij = 0 , (5)
 

where L∗ Jij = c in each case is the likelihood obtained by optimizing all the other parameters with the con-
 
straint of Jij = c . It was proven that 2D obey the χ 2 distribution in a large sample limit (Wilks’ theorem)43.
Accordingly, we may reject the null hypothesis if 2D > zα , where zα is the threshold of χ 2 distribution of a
significance level α. Here we have adopted α = 10−4.

Model validation. The performance of CoNNECT was evaluated using the synthetic data generated by
independent simulations. The presence or absence of connectivity in each direction is decided by whether or
not an output value z ∈ [0, 1] exceeds a threshold θ . It is possible to reduce the number of FPs by shifting the
threshold θ to a high level. But this operation may produce many FNs, making many existing connections be
missed. To balance the false-positives and false-negatives, we considered maximizing the Matthews correlation
coefficient (MCC)44, as has been done in our previous ­study14. The MCC is defined as
NTP NTN − NFP NFN
MCC = √ ,
(NTP + NFP )(NTP + NFN )(NTN + NFP )(NTN + NFN )
where NTP , NTN , NFP , and NFN represent the numbers of true positive, true negative, false positive, and false
negative connections, respectively.
We have obtained two coefficients for excitatory and inhibitory categories and taken the macro-average MCC
that gives equal importance to these categories (Macro-average)45, MCC = (MCCE + MCCI )/2 as we have done
in the previous s­ tudy14. In computing the coefficient for the excitatory category MCCE , we classify connections
as excitatory or other (unconnected and inhibitory); for the inhibitory category MCCI, we classify connections
as inhibitory or other (unconnected and excitatory). Here we evaluate MCCE by considering only excitatory
connections of reasonable strength (EPSP > 0.1 mV for the MAT simulation and > 1 mV for the HH simulation).
We have confirmed that the Matthews correlation coefficient exhibits a wide peak at about θ ∼ 0.5 (Fig. 10),
and accordingly, we adopted θ = 0.5 as the threshold.

A large‑scale simulation of a network of MAT neurons. To obtain a large number of spike trains that
have resulted under the influence of synaptic connections between neurons, we ran a numerical simulation of a
network of 1000 model neurons interacting through fixed synapses. Of them, 800 excitatory neurons innervate
to 12.5 % of other neurons with EPSPs that are log-normally ­distributed14,46–49, whereas 200 inhibitory neurons
innervate randomly to 25% of other neurons with IPSPs that are normally distributed.

Neuron model. As for the spiking neuron model, we adopted the MAT model, which is superior to the Hodg-
kin–Huxley model in reproducing and predicting spike times of real biological neurons in response to fluctuat-
ing ­inputs21,23. In addition, its numerical simulation is stable and fast. The membrane potential of each neuron
obeys a simple relaxation equation following the input signal:
dvm
(6)
 
τm = −(vm − VL ) − τm ge (vm − VE ) + gi (vm − VI ) − RIbg
dt
where ge , gi represents the excitatory conductance and the inhibitory conductance, respectively. Here RIbg rep-
resent the background noise. The conductance evolves with the
dgx gx   
=− + Gj δ t − tjk − dj (7)
dt τs,X j k

Scientific Reports | (2021) 11:12087 | https://doi.org/10.1038/s41598-021-91244-w 13

Vol.:(0123456789)
www.nature.com/scientificreports/

Matthews Correlation Coefficient (MCC)


0.9

0.8

0.7

0.6

0.5

0.4

0.3

0.2

0.1

0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
threshold θ

Figure 10.  The Matthews correlation coefficient (MCC) plotted against the threshold θ for determining the
presence and absence of the connection.

Membrane dynamics
excitatory
τm
inhibitory
, τm (ms) 20, 10
VL , VE, VI (mV) − 70, 0, − 80
τs,e , τs,i (ms) 1, 2
Threshold dynamics
τ1, τ2 (ms) 10, 200
ωexcitatory , ωinhibitory (mV) − 55, − 57
excitatory excitatory
α1 , α2 (mV) Gauss(1.5, 0.25), 0.5
inhibitory inhibitory
α1 , α2 (mV) 3, 0

Table 4.  Parameters for neuron models.

where τs,X is the synaptic time constant, X stands for e (excitatory) or i (inhibitory), tjk is the kth spike time of
jth neuron, dj is a synaptic delay and Gj is the synaptic weight from jth neuron. δ(t) is the Dirac delta function.
Next, the adaptive threshold of each neuron θ(t) obeys the following equation:

θ(t) = H(t − tj ) + ω
(8)
j

 
 t
H(t) = αk exp − (9)
τk
k=1,2

where tj is the jth spike time of a neuron, ω is the resting value of the threshold, τk is the kth time constant, and
αk is the weight of the kth component. The parameter values are summarized in Table 4.

Synaptic connections. We ran a simulation of a network consisting of 800 pyramidal neurons and 200 interneu-
rons interconnected with a fixed strength. Each neuron receives 100 excitatory inputs randomly selected from
800 pyramidal neurons and 50 inhibitory inputs selected from 200 interneurons. The excitatory and inhibitory
synaptic connections were sampled from respective distributions so that the resulting EPSPs and IPSPs are simi-
lar to the distributions adopted in our previous ­study14. In particular, the excitatory conductances {GijE } were
sampled independently from a log-normal ­distribution46,47.
(log x − µ)2
 
1
P(x) = √ exp − , (10)
2πσ x 2σ 2
where µ = −5.543 and σ = 1.30 are the mean and SD of the natural logarithm of the conductances.
The inhibitory conductances {GijI } were sampled from the normal distribution:

Scientific Reports | (2021) 11:12087 | https://doi.org/10.1038/s41598-021-91244-w 14

Vol:.(1234567890)
www.nature.com/scientificreports/

Background
bg bg
ge,0, gi,0 (mS cm−2) 0.123, 0.322
bg bg
Ee , Ei [mV] 0, − 80
bg bg
σe , σi (mS cm−2) 0.0163, 0.0265
bg bg
τs,e , τs,i (ms) 2.7, 10.5

à 0.015

Table 5.  Parameters for synaptic currents and background inputs.

(x − µ)2
 
1
P(x) = √ exp − , (11)
2π σ 2σ 2
where µ = 0.0217 mS cm−2, σ = 0.00171 mS cm−2 are the mean and SD of the conductances. If the sampled
value is less than zero, the conductance is resampled from the same distribution. The delays of the synaptic
connections from excitatory neurons are drawn from a uniform distribution between 3 and 5 ms. The delays of
the synaptic connections from inhibitory neurons are drawn from a uniform distribution between 2 and 4 ms.

Background noise. Because our model network is smaller than real mammalian cortical networks, we added a
background current to represent inputs from many neurons, as previously done by Destexhe et al.11,50.
   
bg bg bg bg
RIbg = ge (t) vm − Ee + gi (t) vm − Ei . (12)

The summed conductance RIbg represents random bombardments from a number of excitatory and inhibitory
neurons. The dynamics of excitatory or inhibitory conductances can be approximated as a stationary fluctuating
process represented as the Ornstein–Uhlenbeck ­process51,

bg bg bg  bg 2
dgX gX − gX,0  2σX
=− bg
+  bg ξ(t), (13)
dt τ τ
s,X s,X

where gX stands for ge or gi, and ξ(t) is the white Gaussian noise satisfying �ξ(t)� = 0 and �ξ(t)ξ(s)� = δij δ(t − s).
The real biological data has a wide variety of fluctuation, including non-trivial large variations with some
characteristic timescales. For instance, the hippocampal neurons are subject to the theta oscillation of the fre-
quency range of 3 − 10 (Hz)52. To reproduce such oscillations that are also observed in the cross-correlogram,
we introduced slow oscillations into the background noise for excitatory neurons, as

bg bg bg  bg 2
dge ge − ge,0  2σe
=− +  bg ξ1 (t) + A sin(ωt + δ)ξ2 (t), (14)
dt bg
τs,e τs,e
where ξ1 (t) and ξ2 (t) are the white Gaussian noise satisfying �ξi (t)� = 0 and �ξi (t)ξj (s)� = δij δ(t − s).
Among N = 1000 neurons, we added such oscillating background signals to three subgroups of 100 neurons
(80 excitatory and 20 inhibitory neurons), respectively with 7, 10, and 20 Hz. The phases of the oscillation δ were
chosen randomly from the uniform distribution. Amplitudes of the oscillations were chosen randomly from
uniform distribution in an interval [Ã/2, 3Ã/2]. The parameters for the background inputs are summarized in
Table 5.

Numerical simulation. Simulation codes were written in C++ and parallelized with OpenMP framework. The
time step was 0.1 ms. The neural activity was simulated up to 7200 s.

Experimental data. Spike trains were recorded from the PF, IT, and V1 cortices of monkeys in three exper-
imental laboratories using the Utah arrays. All the studies were carried out in compliance with the ARRIVE
guidelines. Individual experimental settings are summarized as follows.

Prefrontal cortex (PF). The experiments were carried out on an adult male rhesus macaque Macaca mulatta
(6.7 kg, age 4.5 y). The monkey had access to food 24 h a day and earned liquid through task performance on
testing days. Experimental monkeys were socially pair housed. All experimental procedures were performed in
accordance with the ILAR Guide for the Care and Use of Laboratory Animals and were approved by the Animal
Care and Use Committee of the National Institute of Mental Health (U.S.A.). Procedures adhered to applicable
United States federal and local laws, including the Animal Welfare Act (1990 revision) and applicable Regula-
tions (PL89544; USDA 1985) and Public Health Service Policy (PHS2002). Eight 96–electrode arrays (Utah
arrays, 10 × 10 arrangement, 400 μm pitch, 1.5 mm depth, Blackrock Microsystems, Salt Lake City, U.S.A.) were
­ rocedures53. Briefly, a single bone
implanted on the prefrontal cortex following previously described surgical p

Scientific Reports | (2021) 11:12087 | https://doi.org/10.1038/s41598-021-91244-w 15

Vol.:(0123456789)
www.nature.com/scientificreports/

flap was temporarily removed from the skull to expose the PFC, then the dura mater was cut open to insert the
electrode arrays into the cortical parenchyma. Next, the dura mater was closed and the bone flap was placed back
into place and attached with absorbable suture, thus protecting the brain and the implanted arrays. In parallel,
a custom-designed connector holder, 3D-printed using biocompatible material, was implanted onto the poste-
rior portion of the skull. Recordings were made using the Grapevine System (Ripple, Salt Lake City, USA). Two
Neural Interface Processors (NIPs) made up the recording system, one NIP (384 channels each) was connected
to the 4 multielectrode arrays of one hemisphere. Synchronizing behavioral codes from MonkeyLogic and eye-
tracking signals were split and sent to each NIP. The raw extracellular signal was high-pass filtered (1 kHz cutoff)
and digitized (30 kHz) to acquire single-unit activity. Spikes were detected online and the waveforms (snippets)
were stored using the Trellis package (Grapevine). Single units were manually sorted offline using custom Mat-
lab scripts to define time-amplitude windows in combination with clustering methods based on PCA feature
extraction. Further details about the experiment can be found e­ lsewhere54. Briefly, the recordings were carried
out while the animals were comfortably seated in front of a computer screen, performing left or right saccadic
eye movements. Each trial started with the presentation of a fixation dot on the center of the screen and the
monkeys were required to fixate. After a variable time (400–800 ms) had elapsed, the fixation dot was toggled off
and a cue (white square, 2◦ × 2◦ side) was presented either to the left or right of the fixation dot. The monkeys
had to make a saccade towards the cue and hold for 500 ms. 70% of the correctly performed trials were rewarded
stochastically with a drop of juice (daily total 175–225 mL). Typically, monkeys performed > 1000 correct trials
in a given recording session for recording time of 120-150 min.

Inferior temporal cortex (IT). The experiments were carried out on an adult male Japanese monkey (Macaca
fuscata, 11 kg, age 13 y). The monkey had access to food 24 h a day and earned its liquid during and addition-
ally after neural recording experiments on testing days. The monkey was housed in one of adjoining individual
primate cages that allowed social interaction. All experimental procedures were approved by the Animal Care
and Use Committee of the National Institute of Advanced Industrial Science and Technology (Japan) and were
implemented in accordance with the “Guide for the Care and Use of Laboratory Animals” (eighth ed., National
Research Council of the National Academies). Four 96 microelectrode arrays (Utah arrays, 10 × 10 layout, 400
μm pitch, 1.5 mm depth, Blackrock Microsystems, Salt Lake City, USA) were surgically implanted on the IT
cortex of the left hemisphere. Three arrays were located in area TE and the remaining one in area TEO. Surgical
procedures were roughly the same as having been described p ­ reviously53, except that a bone flap that was tempo-
rarily removed from the skull was located over the IT cortex and that a CILUX chamber was implanted onto the
anterior part of the skull protecting connectors of the arrays. Recordings of neural data and eye positions were
done in a single session using CerebusTM system (Blackrock Microsystems). The extracellular signal was band-
pass filtered (250–7.5 k Hz) and digitized (30 kHz). Units were sorted online before the recording session for
the extracellular signal of each electrode using a threshold and time-amplitude windows. Both the spike times
and the waveforms (10 and 38 samples, preceding and after a threshold crossing, respectively) of the units were
stored using Cerebus Central Suite (Blackrock Microsystems). Single units were refined offline by hand using the
PCA projection of the spike waveforms in Offline sorterTM (Plexon Inc., Dallas, USA). The monkey seated in a
primate chair, and the head was restrained with a head holding device so that the eyes were positioned 57 cm in
front of a color monitor’s display (GDM-F520, SONY, Japan). The display subtended a visual angle of 40◦ × 30◦
with a resolution of 800 × 600 pixels. A television series on animals (NHK’s Darwin’s Amazing Animals, Asahi
Shimbun Publications Inc., Japan) was shown on the display throughout the online spike sorting and the record-
ing session. The monkey’s eye position was monitored using an infrared pupil-position monitoring s­ ystem55 and
was not restricted.

The primary visual cortex (V1). The data set was obtained from Collaborative Research in Computational Neu-
roscience (CRCNS), pvc-1156 by the courtesy of the authors o ­ f57. In this experiment, spontaneous activity was
measured from the primary visual cortex while a monkey viewed a CRT monitor (1024 × 768 pixels, 100 Hz
refresh) displaying a uniform gray screen (luminance of roughly 40 cd/m 2 ). Briefly, the animal was premedi-
cated with atropine sulfate (0.05 mg/kg) and diazepam (Valium, 1.5 mg/kg) 30 min before inducing anesthesia
with ketamine HCl (10.0 mg/kg). Anesthesia was maintained throughout the experiment by a continuous intra-
venous infusion of sufentanil citrate. To minimize eye movements, the animal was paralyzed with a continuous
intravenous infusion of vecuronium bromide (0.1 mg/kg/h). Vital signs (EEG, ECG, blood pressure, end-tidal
PCO2, temperature, and lung pressure) were monitored continuously. The pupils were dilated with topical atro-
pine and the corneas protected with gas-permeable hard contact lenses. Supplementary lenses were used to
bring the retinal image into focus by direct ophthalmoscopy and later adjusted the refraction further to optimize
the response of recorded units. Experiments typically lasted 4–5 d. All experimental procedures complied with
guidelines approved by the Albert Einstein College of Medicine of Yeshiva University and New York University
Animal Welfare Committees.
Spike sorting and analysis criteria: Waveform segments were sorted off-line with an automated sorting algo-
rithm, which clustered similarly shaped waveforms using a competitive mixture decomposition m ­ ethod58. The
output of this algorithm was refined by hand with custom time-amplitude window discrimination software (writ-
ten in MATLAB; MathWorks) for each electrode, taking into account the waveform shape and interspike interval
distribution. To quantify the quality of the recording, the signal-to-noise ratio (SNR) of each candidate unit was
computed as the ratio of the average waveform amplitude to the SD of the waveform n ­ oise59–61. Candidates that
fell below an SNR of 2.75 were discarded as multiunit recordings.

Scientific Reports | (2021) 11:12087 | https://doi.org/10.1038/s41598-021-91244-w 16

Vol:.(1234567890)
www.nature.com/scientificreports/

Received: 25 January 2021; Accepted: 21 May 2021

References
1. Perkel, D. H., Gerstein, G. L. & Moore, G. P. Neuronal spike trains and stochastic point processes: II. Simultaneous spike trains.
Biophys. J. 7, 419 (1967).
2. Toyama, K., Kimura, M. & Tanaka, K. Organization of cat visual cortex as investigated by cross-correlation technique. J. Neuro‑
physiol. 46, 202 (1981).
3. Grun, S. Data-driven significance estimation for precise spike correlation. J. Neurophysiol. 101, 1126 (2009).
4. Amarasingham, A., Harrison, M. T., Hatsopoulos, N. G. & Geman, S. Conditional modeling and the jitter method of spike resa-
mpling. J. Neurophysiol. 107, 517 (2012).
5. Schwindel, C. D., Ali, K., McNaughton, B. L. & Tatsuno, M. Long-term recordings improve the detection of weak excitatory-
excitatory connections in rat prefrontal cortex. J. Neurosci. 34, 5454 (2014).
6. Platkiewicz, J., Saccomano, Z., McKenzie, S., English, D. & Amarasingham, A. Monosynaptic inference via finely-timed spikes. J.
Comput. Neurosci.https://​doi.​org/​10.​1007/​s10827-​020-​00770-5 (2021).
7. Okatan, M., Wilson, M. A. & Brown, E. N. Analyzing functional connectivity using a network likelihood model of ensemble neural
spiking activity. Neural Comput. 17, 1927 (2005).
8. Pillow, J. W. et al. Spatio-temporal correlations and visual signalling in a complete neuronal population. Nature 454, 995 (2008).
9. Stevenson, I. H., Rebesco, J. M., Miller, L. E. & Körding, K. P. Inferring functional connections between neurons. Curr. Opin.
Neurobiol. 18, 582 (2008).
10. Chen, Z., Putrino, D. F., Ghosh, S., Barbieri, R. & Brown, E. N. Statistical inference for assessing functional connectivity of neuronal
ensembles with sparse spiking data. IEEE Trans. Neural Syst. Rehabil. Eng. 19, 121 (2011).
11. Kobayashi, R. & Kitano, K. Impact of network topology on inference of synaptic connectivity from multi-neuronal spike data
simulated by a large-scale cortical network model. J. Comput. Neurosci. 35, 109 (2013).
12. Zaytsev, Y. V., Morrison, A. & Deger, M. Reconstruction of recurrent synaptic connectivity of thousands of neurons from simulated
spiking activity. J. Comput. Neurosci. 39, 77 (2015).
13. Ren, N., Ito, S., Hafizi, H., Beggs, J. M. & Stevenson, I. H. Model-based detection of putative synaptic connections from spike
recordings with latency and type constraints. J. Neurophysiol. 124, 1588 (2020).
14. Kobayashi, R. et al. Reconstructing neuronal circuitry from parallel spike trains. Nat. Commun. 10, 1 (2019).
15. Fukushima, K. Neocognitron: a hierarchical neural network capable of visual pattern recognition. Neural Netw. 1, 119 (1988).
16. LeCun, Y. et al. Convolutional networks for images, speech, and time series. Handb. Brain Theory Neural Netw. 3361, 1995 (1995).
17. LeCun, Y., Bengio, Y. & Hinton, G. Deep learning. Nature 521, 436 (2015).
18. Krizhevsky, A., Sutskever, I. & Hinton, G. E. Imagenet classification with deep convolutional neural networks. Commun. ACM 60,
84 (2017).
19. Aertsen, A. M. & Gerstein, G. L. Evaluation of neuronal connectivity: sensitivity of cross-correlation. Brain Res. 340, 341 (1985).
20. Reid, R. C. & Alonso, J.-M. Specificity of monosynaptic connections from thalamus to visual cortex. Nature 378, 281 (1995).
21. Kobayashi, R., Tsubo, Y. & Shinomoto, S. Made-to-order spiking neuron model equipped with a multi-timescale adaptive threshold.
Front. Comput. Neurosci. 3, 9 (2009).
22. Gerstner, W. & Naud, R. How good are neuron models?. Science 326, 379 (2009).
23. Omura, Y., Carvalho, M. M., Inokuchi, K. & Fukai, T. A lognormal recurrent network model for burst generation during hip-
pocampal sharp waves. J. Neurosci. 35, 14585 (2015).
24. Kobayashi, R. & Kitano, K. Impact of slow K+ currents on spike generation can be described by an adaptive threshold model. J.
Comput. Neurosci. 40, 347 (2016).
25. Barta, T. & Kostal, L. The effect of inhibition on rate code efficiency indicators. PLoS Computat. Biol. 15, e1007545 (2019).
26. Harish, O. & Hansel, D. Asynchronous rate chaos in spiking neuronal circuits. PLoS Comput. Biol. 11, e1004266 (2015).
27. Shinomoto, S., Shima, K. & Tanji, J. Differences in spiking patterns among cortical neurons. Neural Comput. 15, 2823 (2003).
28. Mochizuki, Y. et al. Similarity in neuronal firing regimes across mammalian species. J. Neurosci. 36, 5736 (2016).
29. Stevenson, I. H. Omitted variable bias in GLMs of neural spiking activity. Neural Comput. 30, 3227 (2018).
30. Baker, C., Froudarakis, E., Yatsenko, D., Tolias, A. S. & Rosenbaum, R. Inference of synaptic connectivity and external variability
in neural microcircuits. J. Comput. Neurosci. 48, 123 (2020).
31. Das, A. & Fiete, I. R. Systematic errors in connectivity inferred from activity in strongly recurrent networks. Nat. Neurosci. 23,
1286 (2020).
32. Pillow, J. W., Shlens, J., Chichilnisky, E. & Simoncelli, E. P. A model-based spike sorting algorithm for removing correlation artifacts
in multi-neuron recordings. PLoS ONE 8, e62123 (2013).
33. Shorten, C. & Khoshgoftaar, T. M. A survey on image data augmentation for deep learning. J. Big Data 6, 60 (2019).
34. Wong, S. C., Gatt, A., Stamatescu, V. & McDonnell, M. D. Understanding data augmentation for classification: when to warp? In
2016 International Conference on Digital Image Computing: Techniques and Applications (DICTA) 1–6 (IEEE, 2016).
35. Park, D. S., Chan, W., Zhang, Y., Chiu, C.-C., Zoph, B., Cubuk, E. D. & Le, Q. V. Specaugment: A Simple Data Augmentation
Method for Automatic Speech Recognition. arXiv preprint arXiv:​1904.​08779 (2019).
36. McDonnell, M. D. & Gao, W. Acoustic scene classification using deep residual networks with late fusion of separated high and
low frequency paths. In ICASSP 2020-2020 IEEE International Conference on Acoustics, Speech and Signal Processing (ICASSP)
141–145 (IEEE, 2020).
37. Wang, J. & Perez, L. The effectiveness of data augmentation in image classification using deep learning. Convolut. Neural Netwo.
Vis. Recognit. 11 (2017).
38. Frid-Adar, M. et al. GAN-based synthetic medical image augmentation for increased CNN performance in liver lesion classifica-
tion. Neurocomputing 321, 321 (2018).
39. Cubuk, E. D., Zoph, B., Mane, D., Vasudevan, V. & Le, Q. V. Autoaugment: Learning augmentation strategies from data. In Proceed‑
ings of the IEEE Conference on Computer Vision and Pattern Recognition 113–123 (2019).
40. Kingma, D. P. & Ba, J. Adam: A Method for Stochastic Optimization. arXiv preprint arXiv:​1412.​6980 (2014).
41. Kass, R. E., Eden, U. T. & Brown, E. N. Analysis of neural data, Vol. 491 (Springer, 2014).
42. Volgushev, M., Ilin, V. & Stevenson, I. H. Identifying and tracking simulated synaptic inputs from neuronal firing: insights from
in vitro experiments. PLoS Comput. Biol. 11, e1004167 (2015).
43. Wilks, S. S. The large-sample distribution of the likelihood ratio for testing composite hypotheses. Ann. Math. Stat. 9, 60 (1938).
44. Matthews, B. W. Comparison of the predicted and observed secondary structure of t4 phage lysozyme. Biochim. Biophys. Acta 405,
442 (1975).
45. Sun, A., & Lim, E.-P. Hierarchical text classification and evaluation. In Proceedings of ICDM 2001 521–528 ( IEEE, 2001).
46. Song, S., Sjöström, P. J., Reigl, M., Nelson, S. & Chklovskii, D. B. Highly nonrandom features of synaptic connectivity in local
cortical circuits. PLoS Biol. 3, e68 (2005).
47. Teramae, J.-N., Tsubo, Y. & Fukai, T. Optimal spike-based communication in excitable networks with strong-sparse and weak-dense
links. Sci. Rep. 2, 485 (2012).

Scientific Reports | (2021) 11:12087 | https://doi.org/10.1038/s41598-021-91244-w 17

Vol.:(0123456789)
www.nature.com/scientificreports/

48. Buzsáki, G. & Mizuseki, K. The log-dynamic brain: how skewed distributions affect network operations. Nat. Rev. Neurosci. 15,
264 (2014).
49. Uzan, H., Sardi, S., Goldental, A., Vardi, R. & Kanter, I. Stationary log-normal distribution of weights stems from spontaneous
ordering in adaptive node networks. Sci. Rep. 8, 1 (2018).
50. Destexhe, A., Rudolph, M., Fellous, J.-M. & Sejnowski, T. J. Fluctuating synaptic conductances recreate in vivo-like activity in
neocortical neurons. Neuroscience 107, 13 (2001).
51. Tuckwell, H. C. Introduction to Theoretical Neurobiology: Volume 2, Nonlinear and Stochastic Theories (Cambridge University Press,
Cambridge, 1988).
52. Goutagny, R., Jackson, J. & Williams, S. Self-generated theta oscillations in the hippocampus. Nat. Neurosci. 12, 1491 (2009).
53. Mitz, A. R. et al. High channel count single-unit recordings from nonhuman primate frontal cortex. J. Neurosci. Methods 289, 39
(2017).
54. Bartolo, R., Saunders, R. C., Mitz, A. R. & Averbeck, B. B. Information-limiting correlations in large neural populations. J. Neurosci.
40, 1668 (2020).
55. Matsuda, K., Nagami, T., Sugase, Y., Takemura, A. & Kawano, K. A widely applicable real-time mono/binocular eye tracking system
using a high frame-rate digital camera. In International Conference on Human-Computer Interaction 593–608 (Springer, 2017).
56. Kohn, A., Smith, M. A. Utah array extracellular recordings of spontaneous and visually evoked activity from anesthetized macaque
primary visual cortex (V1). CRCNS.org. http://​dx.​doi.​org/​10.​6080/​K0NC5​Z4X (2016).
57. Smith, M. A. & Kohn, A. Spatial and temporal scales of neuronal correlation in primary visual cortex. J. Neurosci. 28, 12591 (2008).
58. Shoham, S., Fellows, M. R. & Normann, R. A. Robust, automatic spike sorting using mixtures of multivariate t-distributions. J.
Neurosci. Methods 127, 111 (2003).
59. Nordhausen, C. T., Maynard, E. M. & Normann, R. A. Single unit recording capabilities of a 100 microelectrode array. Brain Res.
726, 129 (1996).
60. Suner, S., Fellows, M. R., Vargas-Irwin, C., Nakata, G. K. & Donoghue, J. P. Reliability of signals from a chronically implanted,
silicon-based electrode array in non-human primate primary motor cortex. IEEE Trans. Neural Syst. Rehabil. Eng. 13, 524 (2005).
61. Kelly, R. C. et al. Comparison of recordings from microelectrode arrays and single electrodes in the visual cortex. J. Neurosci. 27,
261 (2007).

Acknowledgements
We thank Masahiro Naito for his technical assistance in developing a web-application program, Jun-nosuke Ter-
amae for his advice on MAT simulation, and Kai Shinomoto for drawing an illustration of a monkey for Figure 1.
We also thank Adam Kohn for permitting us to analyze their experimental data of V1 and providing the detailed
information of the experimental conditions, and Richard Saunders and Mark Eldridge for performing surgery
on the animal for IT cortex data, Yuji Nagai and Takafumi Minamimoto for assisting the surgery, and Rossella
Falcone and Narihisa Matsumoto for helpful discussions upon preparing the IT cortex data. R.K. is supported
by JSPS KAKENHI Grant Numbers JP17H03279, JP18K11560, JP19H01133, and JPJSBP120202201, and JST
PRESTO Grant Number JPMJPR1925, Japan. B.B.A. is supported by NIMH DIRP ZIA MH002928. Y.S.M is
supported by JSPS KAKENHI Grant Number JP18H05020 and New Energy and Industrial Technology Develop-
ment Organization (NEDO). K.H. is supported by Japan Society for the Promotion of Science (JSPS) and JSPS
KAKENHI Grant Number JP19J40302. K.K. is supported by JSPS KAKENHI Grant Number JP19K07804. B.J.R.
is supported by NIMH DIRP ZIA MH002032. S.S. is supported by JST CREST Grant Number JPMJCR1304, and
the New Energy and Industrial Technology Development Organization (NEDO).

Author contributions
D.E. and R.K. contributed equally in performing analysis. R.B., B.B.A., Y.S-M, K.H., and K.K. contributed data.
B.J.R. and S.S. wrote the main manuscript text. S.S. designed the study. All authors discussed the results and
reviewed the manuscript.

Competing interests
The authors declare no competing interests.

Additional information
Correspondence and requests for materials should be addressed to S.S.
Reprints and permissions information is available at www.nature.com/reprints.
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional affiliations.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International
License, which permits use, sharing, adaptation, distribution and reproduction in any medium or
format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the
Creative Commons licence, and indicate if changes were made. The images or other third party material in this
article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the
material. If material is not included in the article’s Creative Commons licence and your intended use is not
permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder. To view a copy of this licence, visit http://​creat​iveco​mmons.​org/​licen​ses/​by/4.​0/.

© The Author(s) 2021

Scientific Reports | (2021) 11:12087 | https://doi.org/10.1038/s41598-021-91244-w 18

Vol:.(1234567890)

You might also like