background image

 

Research Articles | Systems/Circuits 

 

 
Basic Properties of Coordinated Neuronal Ensembles in
the Auditory Thalamus

 

 

https://doi.org/10.1523/JNEUROSCI.1729-23.2024

 
Received: 13 September 2023
Revised: 4 March 2024
Accepted: 11 March 2024
 

Copyright © 2024 the authors

This Early Release article has been peer reviewed and accepted, but has not been through
the composition and copyediting processes.The final version may differ slightly in style or
formatting and will contain links to any extended data.

Alerts: Sign up at 

www.jneurosci.org/alerts

 to receive customized email alerts when the fully

             formatted version of this article is published.

JNEUROSCI.1729-23.2024.full-html.html
background image

Title: Basic Properties of Coordinated Neuronal Ensembles in the Auditory Thalamus

 

Abbreviated title: cNEs in MGB

 

 

Congcong Hu

1,2,3

, Andrea R. Hasenstaub

1,2,3

, Christoph E. Schreiner

1,2,3

*

 

1

John & Edward Coleman Memorial Laboratory, 

2

Neuroscience Graduate Program, 

3

Department 

of Otolaryngology-Head and Neck Surgery, University of California-San Francisco, San 

Francisco, California 94158, USA

 

*

Corresponding author: Christoph E. Schreiner, christoph.schreiner@ucsf.edu

 

 

Number of pages: 51 

10 

Number of figures: 9 

11 

Number of all Words: 13940 

12 

13 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

Abstract (187), Significance Statement (115), Introduction (631), Discussion (1456) 

14 

Conflict of interest statement: The authors declare no competing financial interests.

 

15 

Acknowledgements: We thank Drs. Jermyn See and Natsumi Homma for their help with 

16 

software and data collection.  This work was supported by the National Institutes of Health 

17 

(DC002260 and DC017396 to C.E.S., and DC014101, NS116598, MH122478, and EY025174 to 

18 

A.R.H.), the Klingenstein Foundation (A.R.H.), PBBR Breakthrough Fund (A.R.H.), the 

19 

Coleman Memorial Fund (A.R.H., C.E.S.), and Hearing Research Inc. (C.E.S., A.R.H). 

20 

 

 

21 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

Abstract  

23 

Coordinated neuronal activity has been identified to play an important role in information 

24 

processing and transmission in the brain. However, current research predominantly focuses on 

25 

understanding the properties and functions of neuronal coordination in hippocampal and cortical 

26 

areas, leaving subcortical regions relatively unexplored. In this study, we use single-unit 

27 

recordings in female Sprague-Dawley rats to investigate the properties and functions of groups 

28 

of neurons exhibiting coordinated activity in the auditory thalamus -- the medial geniculate body 

29 

(MGB). We reliably identify coordinated neuronal ensembles (cNEs), which are groups of 

30 

neurons that fire synchronously, in the MGB. cNEs are shown not to be the result of false 

31 

positive detections or byproducts of slow state oscillations in anesthetized animals. We 

32 

demonstrate that cNEs in the MGB have enhanced information encoding properties over 

33 

individual neurons. Their neuronal composition is stable between spontaneous and evoked 

34 

activity, suggesting limited stimulus-induced ensemble dynamics. These MGB cNE properties 

35 

are similar to what is observed for cNEs in the primary auditory cortex (A1), suggesting that 

36 

ensembles serve as a ubiquitous mechanism for organizing local networks and play a 

37 

fundamental role in sensory processing within the brain. 

38 

 

39 

40 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

Significance Statement  

41 

Temporal coordination of neuronal activity has been widely observed in various cortical areas 

42 

and has been shown to be important for signal processing and information transmission in the 

43 

brain. However, it remains unclear whether neuronal coordination is exclusive to cortical local 

44 

networks or if it also holds significance in subcortical regions. We conducted single-unit 

45 

recordings to investigate coordinated neuronal ensembles (cNEs), which are groups of neurons 

46 

with synchronous firing, in both the auditory thalamus and cortex. We demonstrated the 

47 

existence of cNEs in the auditory thalamus, which have similar properties to cNEs in the 

48 

auditory cortex. This provides evidence that subcortical neuronal coordination can serve as a 

49 

fundamental mechanism for organizing and processing neural signals. 

50 

 

51 

52 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

Introduction 

53 

The function of coordinated neuronal activity in cognitive processes has long been a subject of 

54 

interest in systems neuroscience (Konorski, 1948; Hebb, 1949). Initially, such activity was 

55 

difficult to observe experimentally. However, recent technological advancements in large-scale 

56 

recording, such as two-photon imaging and high-density multi-channel probes, have facilitated 

57 

extensive investigations into the properties and functions of coordinated neuronal firing, 

58 

primarily within the hippocampus and neocortex (Laubach et al., 2000; Baeg et al., 2003; Harris 

59 

et al., 2003; Bizley et al., 2010; Buzsáki, 2010; Bathellier et al., 2012; Oberto et al., 2021; 

60 

Boucly et al., 2022; Domanski et al., 2023). These studies have revealed temporal coordination 

61 

among neurons in several brain areas, shedding light on their potential roles in various cognitive 

62 

processes, such as perception, memory formation, and decision making. Indeed, neuronal 

63 

ensembles have been proposed as the fundamental units for information processing and 

64 

transmission (Buzsáki, 2010; Yuste, 2015). 

65 

In sensory systems in particular, temporal coordination among neurons has been proposed as a 

66 

mechanism to enhance information processing (Kreiter and Singer, 1996; Dan et al., 1998; See et 

67 

al., 2018, 2021) and facilitate communication within and between brain regions (Zandvakili and 

68 

Kohn, 2015; Oberto et al., 2022). Neuronal coordination allows more reliable and specific 

69 

representation of stimuli (See et al., 2018; Yoshida and Ohki, 2020; See et al., 2021; Ebrahimi et 

70 

al., 2022), and considering neuronal coordination allows identification of emergent stimulus 

71 

encoding properties (deCharms and Merzenich, 1996; Shahidi et al., 2019). Additionally, 

72 

elevated coordination in neuronal activity in an output or sender area often precedes activity in a 

73 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

target or receiver area (Zandvakili and Kohn, 2015).  Thus, it is crucial to investigate the 

74 

expression of cNE structure and function at various stages along the sensory pathway. 

75 

In the auditory system, A1 contains pairs of neurons with correlated activity (Brosch and 

76 

Schreiner, 1999; Eggermont, 2000; Atencio and Schreiner, 2013) as well as larger groups of 

77 

neurons with correlated firing. These groups contribute to the representation of auditory stimuli 

78 

(Kreiter and Singer, 1996; Miller and Recanzone, 2009; Ince et al., 2013; See et al., 2018). While 

79 

characteristics of coordinated activity within the cortex have recently been studied (Bathellier et 

80 

al., 2012; Chamberland et al., 2017; See et al., 2018, 2021), our understanding of the 

81 

organization and functional significance of neuronal ensembles in subcortical regions, such as 

82 

the thalamus, remain unknown. The thalamus is of particular interest since it is the gateway and a 

83 

direct intermediary between the peripheral sensory system and the cortex (Winer et al., 2005; 

84 

Smith et al., 2012; Bartlett, 2013).  The auditory thalamus MGB and A1 are highly 

85 

interconnected, with structured connections linking neurons which share similar spectral and 

86 

temporal response properties (Miller et al., 2002; Bartlett and Wang, 2007). Considering the 

87 

strong connections between the thalamus and cortex, investigating the shared characteristics of 

88 

neuronal ensembles in both regions will help us better understand the role these ensembles may 

89 

play in auditory information processing and transmission, and sensory processing in general. 

90 

In this study, we aimed to identify and characterize coordinated neuronal ensembles (cNEs) in 

91 

the MGB. We reliably detected cNEs, defined as groups of neurons exhibiting temporally highly 

92 

coordinated activity, in the MGB. The applied detection method showed robustness, consistently 

93 

identifying cNEs across different time bin sizes. Importantly, we observed a high degree of 

94 

similarity between cNEs derived from spontaneous and evoked activity, suggesting that these 

95 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

ensembles represent functional networks that can operate, to a substantial degree, independently 

96 

of specific sensory stimuli. Furthermore, cNEs in the MGB and A1 shared key characteristics. In 

97 

both structures, spikes associated with cNEs reflect auditory information more reliably than 

98 

random spikes from the same neurons. These findings support the hypothesis that cNEs serve as 

99 

a ubiquitous mechanism for organizing local networks and function as fundamental units for 

100 

sensory processing in the brain. 

101 

Materials and Methods 

102 

Animals 

103 

All experimental procedures were approved by the Institutional Animal Care and Use Committee 

104 

at the University of California, San Francisco (UCSF), and followed the guidelines of the 

105 

National Institute of Health for the care and use of laboratory animals. Twenty-four female 

106 

Sprague-Dawley rats (wild type, 250-350g, 2-4 months; RRID: MGI: 5651135), sourced from 

107 

Charles River, were used in this study. 

108 

Surgery 

109 

The detailed procedures were as described in previous studies (See et al., 2018; Homma et al., 

110 

2020). Briefly, anesthesia was induced with a combination of ketamine (100 mg/kg, Ketathesia, 

111 

HenrySchein) and xylazine (3.33 mg/kg, AnaSed, Akorn), along with atropine (0.54 mg/kg, 

112 

AtroJectSA, HenrySchein), dexamethasone (4 mg/kg, Dexium-SP, Bimeda), and meloxicam (2 

113 

mg/kg, Eloxiject, HenrySchein). Additional doses of ketamine (10-50 mg/kg) and xylazine (0-20 

114 

mg/kg) were given as needed to maintain anesthesia. Local anesthesia was provided using 

115 

lidocaine (Lidoject, 2%, HenrySchein) prior to making incisions. The respiratory rate, heart rate, 

116 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

and depth of anesthesia were continuously monitored, and anesthesia was adjusted as needed. 

117 

The body temperature was monitored and maintained at 37°C using a homeothermic blanket 

118 

system (Harvard Apparatus 55-7020). Lubricant ophthalmic ointment (Artificial Tears, 

119 

HenrySchein) was applied to protect the eyes. A tracheotomy was performed to ensure stable 

120 

breathing during recording. To access the brain, the skin, muscle, skull, and dura over the right 

121 

temporal lobe were removed, and silicone oil (Sigma-Alderich) was applied to cover the cortex. 

122 

A bone rongeur was used to widen the craniotomy window and provide dorsal access to the 

123 

MGB. A cisternal drain was performed to prevent brain swelling. 

124 

Electrophysiology 

125 

The frequency organization of auditory cortex was first mapped using Tungsten electrodes. A1 

126 

was identified as the area with a high to low frequency preference gradient on the rostral-causal 

127 

axis and short latency response to pure tones (Polley et al., 2007). Then, electrophysiological 

128 

recordings were performed using a linear silicon probe with 64 channels (H3, 20µm channel 

129 

distance, Cambridge NeuroTech) in the MGB and a 2-shank probe with 64 channels (H2, 25µm 

130 

channel distance, Cambridge NeuroTech) in A1. The ventral division of the MGB is 

131 

characterized by a low to high frequency gradient on the dorsal-ventral axis (Morel et al., 1987; 

132 

Anderson and Linden, 2011). The probes were inserted using microdrives (David Kopf 

133 

Instruments) at a rate of 25 µm/s to a depth of 4500 to 6000 µm from the surface of the cortex to 

134 

reach MGB (Figure 1A) and 900 to 1300 µm in A1 along the columnar structure (Figure 7A), 

135 

respectively. Extracellular voltage traces were recorded at a sampling rate of 20 kHz with an 

136 

Intan RHD2132 Amplifier system (Intan Technologies). Multi-unit (MU) activities (Figure 1A 

137 

and 7A) were defined as negative peaks crossing 4 standard deviations from the mean in the 

138 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

10 

extracellular voltage trace filtered between 300 and 6000 Hz. Single unit activities were obtained 

139 

by spike sorting using Kilosort 2.5 (Steinmetz et al., 2021; Pachitariu et al., 2023), followed by 

140 

manual curation using Phy (

https://github.com/cortex-lab/phy

). In the manual curation process, 

141 

we visually evaluated individual clusters by examining auto-correlograms, spike waveforms, the 

142 

stability of spike amplitude over time, the persistence of activity over time, the cluster's 

143 

separation from noise in the feature space, and other visual aids provided by phy2 for 

144 

distinguishing single-unit clusters from multi-unit or noise clusters. Subsequently, the identified 

145 

units underwent filtering based on specific criteria: inter-spike interval (ISI) violation within 2ms 

146 

< 1.5%. The majority of single units exhibited an ISI violation lower than 0.25%; peak signal-to-

147 

noise ratios (SNRs) of the waveform > 1.5 (median peak SNR: MGB: 4.52, A1: 3.77); and firing 

148 

rates > 0.1 Hz to eliminate potential multi-units. To assess the reliability of activity of the single 

149 

units across the entire recording duration we obtained the presence ratio. This is the ratio of the 

150 

number of blocks where the unit showed activity and the total number of blocks in a recording 

151 

session. To calculate the presence ratio, the entire recording was divided into 100 equal time 

152 

blocks. The majority of the obtained single units were active in more than 95% of the time 

153 

blocks. This sorting resulted in single units that exhibit low ISI violations, high peak SNR, firing 

154 

rates with a log-normal distribution, and a more consistent presence during recording when 

155 

compared to all clusters generated by Kilosort. To identify oscillatory response in MGB or A1, 

156 

single units on the same electrode were combined to form multi-units (Figure 8).  

157 

Stimuli 

158 

To measure frequency tuning, we presented pure tones with frequencies ranging from 0.5 to 32 

159 

kHz in 0.13 octave steps and sound levels from 0 to 70 dB in 5 dB steps (50ms, 5ms ramps). 

160 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

11 

Each frequency-sound level combination was presented once in a pseudo-random order, with an 

161 

inter-stimulus interval of 250ms. To assess spectrotemporal receptive fields (STRFs), we used a 

162 

15-minute dynamic moving ripple (DMR) (Escabí and Schreiner, 2002). The DMR consisted of 

163 

40 sinusoidal carrier frequencies per octave in the range of 0.5 to 40 kHz, each with a random 

164 

phase. The carriers were slowly modulated (maximum rate of change 

 3Hz) by a 

165 

spectrotemporal envelope with a maximum spectral modulation rate of 4 cycles/octave, a 

166 

maximum temporal modulation rate of 40 cycles/s, and a maximum modulation depth of 40 dB. 

167 

The mean intensity of the DMR was set at 70 dB sound pressure level. We selected DMR as the 

168 

stimulus for analyzing cNEs’ response to sound stimuli, as overt onset response effects for the 

169 

15-minute continuous stimulus is negligible. Additionally, DMR has been observed to reduce 

170 

oscillatory states in the neural population (Miller and Schreiner, 2000). All auditory stimuli were 

171 

generated using MATLAB (MathWorks) and calibrated using a 1/2-inch pressure field 

172 

microphone (Type 4192, Bruel and Kjær). The stimuli were delivered contralaterally from the 

173 

recording site using a closed-field electrostatic speaker (EC1, TuckerDavis Technologies) at a 

174 

sampling rate of 96 kHz. 

175 

Detecting coordinated Neural Ensembles (cNEs) 

176 

To identify groups of neurons that exhibit synchronized co-activation, referred to as “cNEs”, we 

177 

used a method combining principal component analysis (PCA) and independent component 

178 

analysis (ICA) (Lopes-dos-Santos et al., 2013; See et al., 2018). We selected a bin size of 10ms 

179 

as a standard synchronization span because it represents the most appropriate time window to 

180 

capture the synaptic integration window of most cortical neurons (Léger et al., 2005; D’amour 

181 

and Froemke, 2015). First, the individual spike trains of simultaneously recorded neurons were 

182 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

12 

binned and z-scored. Next, the z-scored spike matrix underwent PCA to obtain the eigenvalues 

183 

of the spike train correlation matrix. To determine the number of cNEs, we took eigenvalues to 

184 

be significant if their value exceeded the 99.5

th

 percentile of the Marchenko-Pastur distribution, 

185 

which describes the probability density function of eigenvalues of large rectangular random 

186 

matrices (Marchenko and Pastur, 1967; See et al., 2018) (Figure 2A-ii). We then performed ICA 

187 

(FastICA) on the subspace spanned by the eigenvectors corresponding to the significant 

188 

eigenvalues. The resulting independent components (ICs) represent groups of neurons with 

189 

shared spiking events. The weight of each neuron on an IC indicates the neuron’s contributions 

190 

to the cNE (Figure 2A-iii). As the signs of IC weights were arbitrary, for each IC, the direction 

191 

with the largest absolute weight was rendered positive. The length of each IC was normalized to 

192 

one, making an IC with equal contribution from all neurons have weights of 1/√N, where N was 

193 

the number of neurons in the recording. Neurons with weights over 1/√N were referred to as 

194 

“cNE members” (Oberto et al., 2021) (Figure 2A-iv). 

195 

The strength of the cNE activation at each time point was measured by the similarity between the 

196 

activity of cNE members as well as the cNE pattern, i.e., which neurons in a penetration were 

197 

cNE members. The similarity can be measured as the square of the weighted sum of the z-scored 

198 

spike counts, 

𝑠 = 𝑧

𝑇

𝑤𝑤

𝑇

𝑧 =   𝑧

𝑇

𝑃𝑧

, where 

z

 is the z-scored spike counts of cNE members at 

199 

each time point, 

w

 is the IC weights of cNE members, and the projection matrix 

P

 is the outer 

200 

product of 

w

. To consider only co-activation of multiple cNE members, we set the diagonal of 

201 

the projection matrix to zero and obtained the modified projection matrix 

P*

. cNE activity 

202 

strength was calculated as 

𝑠 =   𝑧

𝑇

𝑃

𝑧

. A null distribution of cNE activity was obtained by 

203 

projecting a circularly shifted spike matrix, where the temporal relationship of neurons was 

204 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

13 

disrupted, to the template matrix (See et al., 2018). This process was iterated 50 times, and the 

205 

threshold of cNE activation was defined as the 99.5th percentile of the null distribution (Figure 

206 

2A-v). The spikes of cNE members within the selected time bin where the cNE was active were 

207 

referred to as “cNE spikes”. 

208 

Matching cNEs across different bin sizes 

209 

We used the correlation between IC weights of all neurons in a penetration to assess the 

210 

similarity of cNE patterns across different synchronization windows (i.e., time bin width: 2, 5, 

211 

10, 20, 40, 80 and 160ms). To visualize the similarity of cNEs identified using 10ms bins to 

212 

those identified using other bin sizes, we calculated the cNEs for the same recording using other 

213 

bin sizes. We displayed the cNEs whose IC weights were best correlated with the 10ms cNE 

214 

(Figure 3A). To measure the variability of cNE identities across bin sizes (Figure 3B, C), we 

215 

matched each cNE to the most similar cNE calculated using reference bins (i.e., when using 

216 

10ms as the reference bin size, to the 10ms cNE which had the best-correlated IC weights). The 

217 

proportion of shared members with a reference cNE was then calculated by dividing the number 

218 

of members in a cNE that were also identified as members in its matching reference cNE by the 

219 

total number of unique members in both cNEs combined. 

220 

The significance of the match was determined based on the null distribution of IC weight 

221 

correlations between matched cNEs. For example, to determine the significance of the 

222 

correlation between the IC weights of a 10ms cNE with its most correlated 160ms cNE, we first 

223 

generated a null distribution of IC weight correlations. We circularly shifted spike trains and then 

224 

applied PCA/ICA to identify sham cNEs using the shuffled spike matrices binned at 160ms, 

225 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

14 

maintaining the same number of cNEs as the original 160ms cNEs. Then we identified the most 

226 

correlated sham 160ms cNE for the 10ms cNE. This process was repeated 1000 times to generate 

227 

the null distribution of correlation values. The significance threshold was set at p < 0.01. 

228 

Assessing stability of cNEs 

229 

To assess the stability of cNEs during and across spontaneous and stimulus-driven activity, we compared 

230 

the cNEs from adjacent recording segments (Figure 4A). 

To match the IC weights of cNEs identified 

231 

from the different recording segments, we used an iterative process that involved selecting cNE 

232 

pairs from the two segments with the highest correlations (Spearman’s r) (Oberto et al., 2021). 

233 

First, we computed the correlations between all possible pairs of cNEs that were generated from 

234 

the two segments. Then, the pair with the highest correlation was set aside, and the same process 

235 

was repeated with the remaining cNEs until all cNEs were paired. If there were any remaining 

236 

cNEs that did not have a match due to a difference in the number of cNEs between the two 

237 

segments, they were left unmatched. 

238 

To generate a null distribution of IC weight correlations between matched cNEs from two 

239 

recording segments (Figure 4D), we circularly shifted spike trains within each activity block. We 

240 

then applied PCA/ICA to identify sham cNEs using the resulting shuffled spike matrices. As the 

241 

shuffling disrupted correlations between neurons, very few eigenvalues exceeded the upper 

242 

bounds of the Marchenko-Pastur distribution. To address this, we maintained the number of 

243 

sham cNEs in the shuffled data equivalent to the number of significant eigenvalues obtained 

244 

from the original spike matrix. The sham cNEs from adjacent activity blocks were then matched 

245 

following the procedure described above. This iterative process was repeated 1000 times to 

246 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

15 

establish a null distribution of IC weight correlations for the matched cNEs. The 99.5th 

247 

percentile of each null distribution was set as the significance threshold. 

248 

False positive detection of cNEs

 

 

249 

To assess the potential for false positive cNE detection, we applied the cNE detection algorithm 

250 

to shuffled data, using the same criteria as those applied to the real dataset. This process was 

251 

repeated 10 times, resulting in an average count of false positive cNEs across the circularly 

252 

shifted data (Figure 4F). Despite conducting 10 iterations, false positive cNEs were not 

253 

consistently identified in neighboring blocks. In cases where a false positive cNE was detected 

254 

(i.e., when an eigenvalue computed from shuffled data exceeded the Marchenko-Pastur 

255 

distribution), we evaluated its stability by measuring the highest correlation of its IC weights 

256 

with those of real cNEs in the adjacent block (Figure 4E). The significance of false positive cNE 

257 

IC weight correlations was determined using the same threshold established for real cNEs. 

258 

STRF analysis 

259 

For analysis, we down-sampled DMR to a resolution of 0.1 octaves in frequency and 5ms in 

260 

time. We used the reverse correlation method to obtain the STRFs of the units (Theunissen et al., 

261 

2000; Escabí and Schreiner, 2002). To derive the STRFs, we averaged the spectrotemporal 

262 

envelopes of the stimulus over a period of 100ms preceding spikes (Figure 1B and 6A). Positive 

263 

(red) values on a STRF indicate that the sound energy at that frequency and time tends to 

264 

increase the firing rate of the unit, while negative (blue) values indicate where the stimulus tends 

265 

to decrease the firing rate of the unit. The frequency corresponding to the highest absolute value 

266 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

16 

of the STRF is considered the best frequency (BF) of the unit (Miller et al., 2002). We also 

267 

determined the peak-to-trough difference (PTD) as a measure of STRF strength. 

268 

A STRF was considered significant if it reliably described a neuron's response to DMR sound. 

269 

To assess the reliability of a STRF, we divided a neuron's spikes into two equal halves and 

270 

generated two corresponding STRFs (STRF A and B) using each half (Qiu et al., 2003). The 

271 

similarity between STRF A and B was computed using Pearson's correlation. This process was 

272 

reiterated 1000 times, and the average STRF similarity across these iterations was used as the 

273 

measure of reliability.  To determine the statistical significance of a STRF's reliability, we 

274 

constructed a null distribution by reversing the neuron's spike train, thus disrupting the temporal 

275 

correlation of neural responses to the stimulus. We considered STRFs with a reliability 

276 

surpassing a z-score of 2.58 to be significant. 

277 

We used mutual information (MI) as the metric to quantify the amount of information we can 

278 

obtain about the stimulus by observing spikes of neurons or cNEs (Atencio and Schreiner, 2008; 

279 

See et al., 2018). The stimulus segment 

𝑠 

preceding each spike was projected onto the STRF via 

280 

the inner product 

𝑧  =  𝑠  ∗  𝑆𝑇𝑅𝐹 

.The projection values were then binned to get the probability 

281 

distribution  

𝑃(𝑧|𝑠𝑝𝑖𝑘𝑒)

. The 

a priori

 distribution of stimulus projection values, 

𝑃(𝑧)

, was 

282 

calculated by projecting all stimulus segments of DMR onto the STRF, regardless of spike 

283 

occurrence. Both distributions 

𝑃(𝑧)

 and 

𝑃(𝑧|𝑠𝑝𝑖𝑘𝑒)

 were normalized relative to the mean 

𝜇

 

 and 

284 

standard deviation 

𝜎

 

 of 

𝑃(𝑧)

, by 

𝑥  =  

(𝑧−𝜇)

𝜎

 , resulting in 

𝑃(𝑥)

 and 

𝑃(𝑥|𝑠𝑝𝑖𝑘𝑒)

. The MI between 

285 

STRF projection values and single spikes was computed according to 

𝐼  =

286 

  ∫ 𝑑𝑥𝑃(𝑥|𝑠𝑝𝑖𝑘𝑒) log

2

𝑃(

𝑥

|

𝑠𝑝𝑖𝑘𝑒

)

𝑃(𝑥)

287 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

17 

STRF comparisons between cNEs and non-cNE groups of neurons 

288 

To control for the potential influence of population synchrony on a cNE due to independent 

289 

neuron activity, we compared STRFs derived from cNEs and non-cNE groups of neurons. If less 

290 

than half of members had significant STRFs, the cNE was excluded from analysis.  

291 

First, we compared the group STRFs of cNEs and non-cNE groups of neurons (Figure 6B-ii and 

292 

C-ii) (See et al., 2018). The group STRF was calculated using all spikes from neurons within a 

293 

group. To generate non-cNE groups of neurons relative to a cNE, we first selected one neuron 

294 

from the cNE and then sampled from the remaining neurons with significant STRFs within a 

295 

penetration, forming a group with the same number of neurons as the cNE for comparison. This 

296 

process excluded other member neurons of the cNE. This procedure was repeated for all 

297 

members of the cNE, generating all possible combinations of neurons, each including exactly 

298 

one member neuron from the cNE under examination. Combinations of neurons that included 

299 

more than one neuron from any other cNEs in the same recording were then also excluded. For 

300 

each cNE/non-cNE group comparison, we subsampled the spikes in the cNE and the non-cNE 

301 

groups to the same number. Subsequently, STRF peak-to-trough difference (PTD) and mutual 

302 

information (MI) of the cNE group were compared to the median values of the non-cNE groups. 

303 

We also compared the STRFs of cNE spikes with those of coincident spikes from a single neuron 

304 

(Figure 6B-iii and C-iii) to assess the influence of random coincidence on stimulus preference. 

305 

To obtain coincident spikes from a specific neuron, we first sampled neurons from the recorded 

306 

population that do not share membership with the neuron under examination in any cNE to 

307 

create a non-cNE group. We kept the number of neurons in the non-cNE group the same as the 

308 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

18 

cNE to which the neuron being examined belongs. This sampling process was restricted to 

309 

neurons exhibiting significant STRFs. The coincident spikes of the cNE member refer to spikes 

310 

within 10ms of spikes from other neurons within the non-cNE group. We repeated this procedure 

311 

to generate all possible combinations of non-cNE groups, each containing the cNE member and 

312 

excluding any neuron that shares membership with the neuron under scrutiny. Coincident spike 

313 

trains with less than 100 events were discarded. For each cNE spike/non-cNE spike comparison, 

314 

we subsampled the cNE spikes and spikes from the non-cNE groups to the same number. 

315 

Subsequently, the cNE spike STRF PTD and MI were compared to the median values of the 

316 

random spike STRF PTD and MI from the non-cNE groups. 

317 

Quantifying slow oscillations in neural activity 

318 

To determine whether the neural activity in a recording showed a prominent pattern of slow 

319 

oscillations, we measured silence density and the coefficient of variation (CV) of MU firing rate. 

320 

Silence density was defined as the fraction of 20ms time bins with no population activity (zero 

321 

spikes) (Mochol et al., 2015). The CV of MU firing rate was calculated as CV = 

𝜎

/

𝜇

 

, where 

𝜇

 

is 

322 

the mean firing rate and 

𝜎

 is the standard deviation of the firing rate binned at 20ms time bins. 

323 

Permutation test 

324 

We used permutation tests to determine the statistical significance of differences in cross-

325 

correlograms (CCGs) among neurons based on their membership (Figure 2C and 3D), as well as 

326 

to assess differences in the proportion of stable cNEs between different stimulus conditions 

327 

(Figure 4E and 7E). For example, to assess the difference in CCGs between member pairs and 

328 

non-member pairs (Figure 2C), we shuffled the membership labels of the CCGs and calculated 

329 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

19 

the difference between the average CCGs of member and non-member pairs. We repeated this 

330 

process 10,000 times to generate null distributions of the CCG difference for each data point. 

331 

The 0.5

th

 and 99.5

th

 percentiles of the null distribution were taken as the cutoffs for significance. 

332 

We considered consecutive time bins around 0ms-lag with p < 0.01 to be significant. To assess 

333 

the difference in the proportion of stable cNEs, we shuffled the stimulus condition label 

334 

(spontaneous [‘spon’], DMR [‘dmr’], or cross condition comparison [‘cross’]) and repeated this 

335 

process 10,000 times to generate a null distribution of the difference in proportion. The 

336 

significance level was then determined based on the null distribution. 

337 

Statistics 

338 

Statistical analyses were performed in Python. To compare two unpaired groups (e.g., Figure 

339 

2B), we used Mann–Whitney U tests. To compare two paired groups (e.g., Figure 4F), we used 

340 

Wilcoxon signed-rank tests. Permutation tests (e.g., Figure 2C), and Monte Carlo methods (e.g., 

341 

Figure 4D) were used as described above. To determine if two samples are drawn from the same 

342 

distribution, we used Kolmogorov-Smirnov test (Figure 3D). The specific applications of these 

343 

tests are explained in the results section and figure legends. Significance levels are noted as n.s. 

344 

(p >= 0.05), * (p 

0.05), ** (p 

0.01) and *** (p 

0.001). 

345 

 

346 

Results 

347 

Auditory responses in MGB 

348 

We conducted extracellular recordings in the rat MGB (Figure 1A) using a 64-channel linear 

349 

probe, which allowed us to cover most of its span along the dorsal-ventral axis. To obtain the 

350 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

20 

tonal response properties of the recording sites, we presented pure tones of various frequencies 

351 

and intensities. In the MGB, we usually observed a gradient in the frequency preference of multi-

352 

unit (MU) responses from low to high along the dorsal-ventral axis, which could vary gradually 

353 

(Figure 1A-i) or abruptly (Figure 1A-ii) depending on the probe’s location. Responses on most 

354 

channels in the tonotopic region exhibited clear frequency tuning (between the red lines in Figure 

355 

1A), which likely reflect activities in the ventral MGB, the primary input station to the A1. We 

356 

included all SUs from the MGB in our analysis after spike sorting, without distinguishing 

357 

between sub-regions although the vast majority was likely from the ventral nucleus according to 

358 

its tonotopic organization.  

359 

To estimate the STRFs of SUs, we used a 15-minute DMR stimulus, which is a broadband noise 

360 

with varying spectral and temporal modulation (Escabí and Schreiner, 2002). The STRFs of 

361 

MGB neurons also showed a clear gradient in frequency preference from low to high along the 

362 

dorsal-ventral axis (Figure 1B), consistent with the MU responses to pure tones. We then 

363 

examined the firing correlations between pairs of simultaneously recorded SUs.  MGB neuron 

364 

pairs showed widely different correlations in their firing activity, even if they were close in 

365 

proximity and had similar STRFs. For example, Neurons #1, #2, and #3 had similar receptive 

366 

fields (Figure 1B). While neurons #1 and #3 showed correlated firing in both stimulus-driven 

367 

and spontaneous activity, neurons #2 and #3 showed no significant correlation in their activity 

368 

despite similar STRFs (Figure 1C).  This diversity of correlation patterns, even among neurons 

369 

with similar receptive fields, parallels what was previously observed in the cortex (Brosch and 

370 

Schreiner, 1999; Eggermont, 2000; Atencio and Scheiner, 2013; See et al., 2018; Mogensen et 

371 

al., 2019; Wahlbom et al., 2021). Although the role of neuronal coordination in information 

372 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

21 

processing in the cortex has been extensively proposed and studied (e.g., Paninski et al., 2004; 

373 

Bizley et al., 2010; Buzsáki, 2010; Carrillo-Reid et al., 2015; See et al., 2018), less is known 

374 

about the organization of neuronal ensembles in subcortical regions. Therefore, we aimed to 

375 

identify clusters of neurons that exhibit consistent synchronized firing in the MGB and compared 

376 

the properties of these ensembles with those in A1. 

377 

Identifying groups of neurons with coordinated firing in MGB 

378 

To identify cNEs, i.e., groups of neurons with synchronous firing, we performed a combined 

379 

principal- and independent-component analysis (PCA-ICA) (Lopes-dos-Santos et al., 2013; See 

380 

et al., 2018). The procedure for detecting cNEs in a population of neurons is demonstrated in 

381 

Figure 2A using a recording of spontaneous activity from the MGB (Figure 2A). Among the 20 

382 

isolated single units in the recording, some pairs of neurons had highly correlated firing with 

383 

each other, as shown in dark red in the correlation matrix, while others showed low correlation 

384 

(Figure 2A-i). We performed PCA on the correlation matrix of 10ms-binned spike trains, 

385 

resulting in 20 eigenvalues and corresponding eigenvectors or principal components (PCs) 

386 

(Figure 2A-ii). These eigenvalues describe the contribution of each PC to the variance in the 

387 

neural population activity. To determine the significance of the patterns extracted by PCA, we 

388 

compared the eigenvalues to a threshold drawn based on the Marchenko-Pastur distribution 

389 

(Peyrache et al., 2010; Lopes-dos-Santos et al., 2013) (Figure 2A-ii). In this example recording, 

390 

we observed four significant eigenvalues above the threshold, indicating the presence of four 

391 

detectable cNEs in the recorded population. Although PCA efficiently extracts ensemble 

392 

patterns, it has some limitations due to its variance maximization framework. When two 

393 

ensembles account for similar variance in the data on their corresponding axis, the first PC will 

394 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

22 

represent the average of the two instead of an individual ensemble. This problem is even more 

395 

pronounced when ensembles share neurons. To overcome these limitations, we applied ICA to 

396 

the subspace spanned by the significant PCs (Figure 2A-iii). This approach is not constrained by 

397 

the orthogonality requirement of PCA, allowing for a more precise identification of individual 

398 

cNEs. After the PCA-ICA procedure, we obtained the weights of neurons on the axes that define 

399 

cNEs in the neural population, which were color-coded as columns in Figure 2A-iii. Neurons 

400 

were considered members of a cNE if their IC weights were higher than what would be expected 

401 

from an even weight contribution from all neurons (Figure 2A-iv). The activity resulting from 

402 

co-activation of cNE members can be obtained by projecting the spike matrix on the 

403 

corresponding IC weights of the cNE. To determine the significance of cNE activity magnitude, 

404 

we generated a null distribution of the cNE activity values by circularly shuffling spike trains 

405 

and set the significance criteria at 99.5% (Figure 2A-v). For example, when cNE #1 was active, 

406 

multiple member neurons (2-5 out of 6) fired together. The combined PCA-ICA approach 

407 

provided a useful framework to investigate the organization and function of coordinated neuronal 

408 

activity in the MGB. 

409 

To provide evidence that cNEs captured groups of neurons with correlated firing, we compared 

410 

the correlations of 10ms-binned spike trains based on their cNE membership. Pairs of neurons 

411 

that participated in the same cNE (“member pairs”) exhibited significantly higher correlations 

412 

compared to pairs of neurons that did not share membership in any cNE (“non-member pairs”) 

413 

(Figure 2B). To examine the correlation between member and non-member pairs at a finer 

414 

timescale, we cross-correlated the spike trains using 1ms bins. The correlation among cNE 

415 

member pairs was significantly higher compared to non-member pairs within the [-50, 40] ms 

416 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

23 

lag window (red-shaded area in Figure 2C, bottom left). These results provided evidence that 

417 

groups of neurons with coordinated firing exist in the MGB, and that their coordination was 

418 

captured by the PCA-ICA procedure, resulting in the identification of cNEs.  

419 

Variability in cNE identity for different bin sizes 

420 

The temporal frame used to identify neuronal ensembles plays a critical role in shaping our 

421 

understanding of their nature and function (Buzsáki, 2010). To investigate how different choices 

422 

of timescale affect the identification of cNEs, we calculated cNEs using various spike train bin 

423 

sizes, ranging from 2ms to 160ms (Figure 3). Some cNEs showed consistent IC weights across 

424 

different time bin sizes, with a high correlation to IC weights obtained using 10ms bins (Figure 

425 

3A-i). Other cNEs could only be consistently identified using smaller time bin sizes but deviated 

426 

from those when assessed with larger bin sizes (Figure 3A-ii). We matched cNEs calculated 

427 

using different bin sizes to evaluate their similarity (Figure 3B). With only small change in the 

428 

bin size, the cNEs identified were highly similar. For example, 96% of cNEs identified using 

429 

10ms bins had a significant match with 20ms cNEs. However, when compared to cNEs 

430 

calculated with larger differences in bin sizes, their identities could vary substantially. For 

431 

example, 43% of 10ms cNEs did not show a significant match with 160ms cNEs (Figure 3B). 

432 

The remaining 57% of 10ms cNEs that significantly matched 160ms cNEs exhibited high 

433 

correlation (> 0.6) in their IC weights (Figure 3C-i). Moreover, the majority of significantly 

434 

matched cNEs shared more than half of their neuron membership. Nonetheless, approximately 

435 

25% of 10ms cNEs had no common members with the 160ms cNEs (Figure 3C-ii). There was no 

436 

significant difference in the firing rate of MGB neurons participating in 10ms and 160ms cNEs 

437 

(Figure 3D). In summary, small variations in time bin sizes have a limited effect on cNE identity. 

438 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

24 

However, using large time bin sizes to identify cNEs, such as 160ms, may result in the loss of 

439 

half or more of the cNEs identified using small bin sizes, such as 10-20ms.  

440 

In cases where differences in cNE membership arise due to different bin sizes, we investigated 

441 

the firing correlations between neurons that had shifted in or out of the ensemble. We compared 

442 

the membership of neurons in cNEs identified using 10ms and 160ms bin sizes (“10ms cNEs” 

443 

and “160ms cNEs”, Figure 3E) and categorized neurons in each cNE as the following: members 

444 

in both 10ms and 160ms cNEs, members only in the 10ms cNE, or members only in the 160ms 

445 

cNE. Neurons sharing memberships in both 10ms and 160ms cNEs (stable members) had 

446 

positive correlations in their firing (Figure 3E-i). Neurons only in 160ms cNEs were positively 

447 

correlated with stable members, although the correlation was significantly weaker in the [-17, 

448 

10] ms window (Figure 3E-ii) compared to the correlation among stable members (Figure 3E-i). 

449 

Furthermore, some members of 10ms cNEs were not identified as members in 160ms cNEs. 

450 

These neurons showed no significant difference in their correlations with stable members (Figure 

451 

3E-iii) compared to the correlation among stable members (Figure 3E-i). Therefore, using wider 

452 

bin sizes to identify cNEs results in neurons with weak correlations being included in the 

453 

ensemble, as well as neurons with strong correlations being omitted.  

454 

Variability in cNE structures across spontaneous and evoked activity 

455 

Several studies have shown that cortical neuronal ensembles have stable structures across 

456 

spontaneous and stimulus-driven activity, suggesting a consistent local network organization 

457 

utilized in processing stimulus information (Jermakowicz et al., 2009; Luczak et al., 2009; See et 

458 

al., 2018; Filipchuk et al., 2022). To investigate if this property also exists in cNEs in the MGB, 

459 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

25 

as was observed in A1 (See et al., 2018), we recorded continuous segments of neural activity in 

460 

the absence of sound (spontaneous, hereafter ‘spon’) and during the presentation of the DMR 

461 

stimulus (‘dmr’). We divided each activity type into two segments and detected cNEs in each 

462 

segment separately. We then compared the stability of the cNEs within and across stimulus 

463 

conditions, measured by the correlation of IC weights between adjacent segments (Figure 4A and 

464 

B). We observed that while some cNEs exhibited high stability across stimulus conditions, with 

465 

IC weights correlation comparable to that within the same stimulus condition (Figure 4C-i), 

466 

others showed structures that were less stable across stimulus conditions compared to within a 

467 

stimulus condition (Figure 4C-ii). Using null distributions generated by circularly shuffling 

468 

spikes, we determined the significance of the IC weight correlations and found that both 

469 

examples in Figure 4C were significantly stable across stimulus conditions, although one was 

470 

slightly more stable than the other (Figure 4D). In MGB, within spontaneous or stimulus-driven 

471 

activity, around 80% of cNEs exhibited stable structures across adjacent activity blocks (Figure 

472 

4E). Significantly fewer cNEs (54.8%) were stable across stimulus conditions than within a 

473 

stimulus condition (spon vs cross, p = 0e-4; dmr vs cross, p = 6e-4; spon vs dmr, p = 1.0, 

474 

permutation test with Bonferroni correction). The results provide evidence for the stability of 

475 

cNEs in the MGB, during both spontaneous and stimulus-driven activity, although fewer cNEs 

476 

exhibit stable structures across different stimulus conditions than within the same stimulus 

477 

condition. 

478 

To test the possibility of false positive detection of cNEs, we generated shuffled data on each 

479 

segment by circularly shifting spike trains to disrupt their temporal correlations. We then applied 

480 

the cNE detection algorithm to the shuffled data using the same criteria as for the real data. Our 

481 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

26 

analysis revealed a drastically lower number of cNEs identified in shuffled segments (real data 

482 

vs. shuffled data, spon1: 3.2 ± 0.9 vs 0.2 ± 0.3; spon2: 3.1 ± 0.7 vs 0.3 ± 0.4; dmr1: 3.1 ± 0.9 vs 

483 

0.1 ± 0.2; dmr2: 3.2 ± 1.1 vs 0.1 ± 0.1, mean ± SD) (Figure 4F), suggesting that the chance of 

484 

false positive detection of cNEs is quite low. Furthermore, any cNEs identified in the shuffled 

485 

data did not exhibit the same stability across stimulus conditions observed in real data (Figure 

486 

4E). In summary, our findings indicate that the detection of cNEs in the MGB is reliable and not 

487 

susceptible to false positives. Moreover, the properties of cNEs we observe, such as their 

488 

stability across stimulus conditions, are genuine and not artifacts of random data. 

489 

cNE properties in MGB 

490 

We determined some basic structural properties of MGB cNEs. The spontaneous activity in 34 

491 

MGB recordings revealed 115 cNEs with 3.4 ± 0.9 cNEs per penetration. More cNEs were 

492 

observed in penetrations that captured a higher number of isolated single units (Figure 5A). The 

493 

mean cNE size was 4.3 ± 1.5 members, dependent on the number of isolated neurons (Figure 

494 

5B). Of the 407 neurons isolated in MGB, the majority (78.6%) belonged to a single cNE, 11.5% 

495 

did not belong to any cNE, and 9.8% belonged to multiple cNEs (Figure 5C).  

496 

Next, we investigated whether cNE members were physically and functionally closer to each 

497 

other than non-member pairs of neurons. The pairwise spatial distance of cNE members was 

498 

significantly smaller than that of non-member pairs of neurons in MGB (Figure 5D-i). Moreover, 

499 

the span of cNEs, defined as the longest pairwise distance among all members, was shorter than 

500 

that of randomly selected groups of neurons in the recording (Figure 5D-ii). The tuning of cNE 

501 

members was also closer to each other, as the difference in the BF between cNE member pairs 

502 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

27 

was smaller than that of non-member pairs (Figure 5E-i). The BF span of cNEs, defined as the 

503 

largest difference in BF among all members, was smaller than that of randomly selected groups 

504 

of neurons (Figure 5E-ii). Our results demonstrate that cNEs in the MGB are composed of 

505 

neurons that are physically and functionally closer to each other than non-member pairs, 

506 

suggesting a pattern of local circuit organization as well as local functional congruence. 

507 

While we observed similarity in the tuning of cNE members (Figure 5E-ii), cNEs are not limited 

508 

to groups of neurons with similar receptive fields. The distribution of the pairwise difference in 

509 

the BF of cNE members exhibits a long tail, where cNE members could differ in their BF by 

510 

more than 4 octaves (Figure 5E-i). This suggests that neurons with largely different tuning 

511 

properties also exhibit synchronous activities and participate in the same cNE.  

512 

Previous studies have proposed a manifestation of ensemble coding based on neuronal groups 

513 

with covarying firing rate (Wills et al., 2005; Niessing and Friedrich, 2010; Aschauer et al., 

514 

2022), e.g., groups of neurons that jointly increased their firing rates in response to various pure 

515 

tones (Aschauer et al., 2022). Assessing coactivation based on firing rate in response to various 

516 

stimulus, however, is a limited basis for the identification of neurons that functionally cooperate. 

517 

Such methods may simply reflect coactivation due to the overlap between receptive fields along 

518 

the tonotopic axis, and distinct groups of neurons may emerge due to best-frequency 

519 

discontinuities in the tonotopic organization (see our Figure 1A; Imaizumi et al., 2004). A more 

520 

stringent criterion, that of tight temporal synchrony, as utilized here, can help differentiating 

521 

between neuron groups based on coincidental coactivation and groups based on synchronous 

522 

coactivation.  

523 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

28 

cNEs enhance stimulus information encoding 

524 

Synchronization of neuronal spikes in the cortex has been found to enhance information 

525 

encoding about a stimulus compared to the participating neurons alone (Dan et al., 1998; Atencio 

526 

and Schreiner, 2013; See et al., 2018). This is consistent with the multiplexed nature of an 

527 

individual spike train, whereby spikes representing distinct stimulus aspects are mixed but can be 

528 

separated based on their synchrony with other neurons (Lankarany et al., 2019; See et al., 2021). 

529 

To investigate whether spikes from individual neurons that participate in cNEs also exhibit 

530 

differential coding compared to the neuron’s entire spike train, we compared the STRFs 

531 

calculated using all the spikes emitted by a neuron to STRFs only based on the subset of spikes 

532 

that contributed to cNE events (‘cNE spikes’; Figure 6A). The spike trains were subsampled to 

533 

ensure an equal number of spikes across conditions. Our analysis revealed that the STRFs of 

534 

cNE spikes exhibit stronger excitatory and inhibitory fields compared to the STRFs of all spikes 

535 

from the same neuron, as evidenced by the larger peak-trough difference (PTD) of the cNE 

536 

STRFs (Figure 6B-i), which quantifies the difference between the largest and smallest value in 

537 

the STRF. Given that PTD only considers two extreme values in the STRF, we further evaluated 

538 

the reliability of cNE spikes relative to all spikes in encoding the sound features represented by 

539 

their STRFs by calculating the MI between the stimulus and the spikes. Our results demonstrate 

540 

that cNE-spike STRFs have higher MI than STRFs constructed from all spikes (Figure 6C-i).  

541 

To demonstrate that the increased information conveyed by cNE spike STRFs was not simply 

542 

because cNEs integrate signals over multiple neurons, thus must enhanced information through 

543 

population encoding, we compared STRFs derived from cNE member and non-member neurons. 

544 

First, we compared the multi-unit STRFs of cNE member neurons (cNE group STRF) with those 

545 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

29 

of a neuronal group devoid of neurons sharing membership from any cNE (non-cNE group 

546 

STRF). The cNE group STRFs exhibited a significantly higher MI than that of the non-cNE 

547 

STRFs, although no significant difference in PTD was observed (Figure 6B-ii and C-ii). As the 

548 

group STRFs did not take spike synchrony into account, we further compared cNE spikes and 

549 

coincident spikes of a cNE member. The coincident spikes refer to spikes that occurred within a 

550 

10ms window relative to firing of other neurons not sharing membership with the neuron under 

551 

examination. Both the PTD and MI of cNE spike STRFs were significantly higher than that of 

552 

the coincident spike STRFs (Figure 6B-iii and C-iii).  

553 

Collectively, these findings suggest that cNE spikes can enhance information processing by 

554 

increasing the signal-to-noise ratio and promoting more consistent encoding of certain stimulus 

555 

features compared to including all spikes from the neuron. Furthermore, the enhanced 

556 

information encoding of cNEs is not a trivial result of population encoding but rather hinges on 

557 

the identity of cNE members and synchronous spike events. 

558 

MGB and A1 cNEs have similar properties 

559 

The properties and functions of cNEs have previously been explored within the primary auditory 

560 

cortex (A1; See et al., 2018, 2021), whereas investigations into cNEs in subcortical regions are 

561 

limited. Hence, we aim to determine whether the properties of cNEs in the MGB differ 

562 

substantially from those observed in A1 cNEs, or if they share similarities. To target A1, we used 

563 

a 2-shank probe with 64 channels. The MU responses to pure tones from the two shanks of the 

564 

probe exhibited similar frequency tuning, as the shanks sampled nearby cortical columns (Figure 

565 

7A). The responses on each shank showed small variation in their frequency preference along the 

566 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

30 

depth of the probe, as neurons in the same cortical column have consistent characteristic 

567 

frequencies across the active middle and deep cortical layers (Atencio and Schreiner, 2010; 

568 

Merzenich et al., 2013). Much like cNEs in MGB, cNE spike STRFs in A1 exhibited higher 

569 

peak-trough differences (PTD) and mutual information (MI) compared to all spike STRFs 

570 

(Figure 7B-i and C-i). Compared to A1 neurons, MGB neurons displayed significantly higher 

571 

STRF PTD (all spike STRF PTD, p = 3.1e-42; cNE spike STRF PTD, p = 8.6e-36) and mutual 

572 

information (all spike STRF MI, p = 1.9e-25; cNE spike STRF MI, p = 4.2e-22). We did not 

573 

observe, however, a significant difference between MGB and A1 cNEs regarding their gain in 

574 

cNE spike STRF PTD and MI values over that of member neuron spiking (Figure 7B-ii and C-

575 

ii).  

576 

Addressing concerns of potential false positive detections in A1, we compared the number of 

577 

cNEs detected on real and shuffled activities. A substantially smaller number of cNEs were 

578 

detected on shuffled data compared to real data (real data vs. shuffled data, spon1: 4.6 ± 1.5 vs 

579 

0.9 ± 0.6; spon2: 4.2 ± 1.2 vs 1.1 ± 0.6, dmr1: 4.3 ± 1.5 vs 0.9 ± 0.6, dmr2: 4.2 ± 1.2 vs 1.0 ± 0.8, 

580 

mean ± SD) (Figure 7D). Moreover, A1 cNEs were mostly stable across stimulus conditions, 

581 

similar to MGB cNEs, whereas false positive cNEs did not show such stability (Figure 7E). The 

582 

similarity between MGB and A1 cNEs in their stability across stimulus conditions and enhanced 

583 

information encoding provides support for the concept of cNEs serving as a universal 

584 

mechanism for neuronal organization and information processing. 

585 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

31 

cNE formation does not rely on strong slow oscillations 

586 

Slow-wave oscillations, characterized by alternating periods of large and sustained network 

587 

activity (UP states) and neural quiescence (DOWN states), are frequently observed in the cortex 

588 

and thalamus during anesthesia (Steriade et al., 1993; Contreras et al., 1996; Sanchez-Vives and 

589 

McCormick, 2000; Hasenstaub et al., 2007; Chauvette et al., 2011; Neske, 2016). To quantify the 

590 

level of slow oscillations in the recording and their potential influence on cNE properties, we 

591 

used two measurements: silence density and the coefficient of variation (CV) of the multiunit 

592 

(MU) firing rate. Silence density represents the proportion of recording time when no spike was 

593 

fired by the neural population, which is characteristic of the DOWN state in slow oscillations. In 

594 

brains without strong slow oscillations, the population of neurons fires continuously, resulting in 

595 

low silence density. The MU firing rate CV measures the level of variation in the firing rate of 

596 

the MU, which is high for neurons going through UP-DOWN state cycles, but small for neural 

597 

populations with less synchronized oscillatory activity. Recordings with high silence density and 

598 

high MU firing rate CV showed prominent slow oscillations, with epochs of synchronous firing 

599 

of neurons and epochs of quiescence with no spikes (Figure 8A-i). In contrast, recordings with 

600 

low silence density and low MU firing rate CV did not exhibit strong slow oscillations, 

601 

displaying relatively stable and continuous firing (Figure 8A-ii). Recordings with moderate 

602 

silence density and MU firing rate CV exhibited moments of elevated firing, although not as 

603 

synchronized as in recordings with strong oscillations (Figure 8A-iii). By utilizing silence 

604 

density and MU firing rate CV, we were able to differentiate recordings without strong slow 

605 

oscillations from those with strong slow oscillations. 

606 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

32 

The relationship between cNEs and slow oscillations was investigated by applying the cNE 

607 

detection algorithm to recordings without strong oscillations in response to DMR during 15-

608 

minute recordings. Recordings exhibiting low silence density and low CV of the MU firing rate 

609 

were selected to ensure the absence of strong slow oscillations (n(animals): MGB = 8, A1 = 18; 

610 

n(recordings): MGB = 13, A1 = 10). To determine whether cNEs were solely a byproduct of 

611 

slow oscillations, we compared the number of cNEs detected in these recordings against the 

612 

expected false positive rate. We found a significantly higher occurrence of cNEs in activity 

613 

characterized by low silence density and low MU firing rate CV in both MGB (n(cNE) = 3.9 ± 

614 

1.0, mean ± SD; p = 2.4e-4, Wilcoxon signed-rank test) and A1 (n(cNE) = 5.2 ± 1.9, p = 0.002) 

615 

when compared to shuffled data (as in Figure 5). This supports the notion that cNEs are not a 

616 

byproduct of slow oscillations. 

617 

Slow oscillations in thalamic and cortical firing rates are commonly observed and can be related 

618 

to synchronized and desynchronized states of the system (Metherate and Ashe, 1993; Steriade et 

619 

al., 1993; Cowan and Wilson, 1994; Sanchez-Vives and McCormick, 2000; Hasenstaub et al., 

620 

2007). However, the effect of firing rate changes on the information carried by cNEs is not 

621 

known. Therefore, we tested whether cNE spikes exhibit enhanced information encoding in 

622 

recordings without strong oscillations in response to the stimulus. The results showed that cNE 

623 

spike STRFs have higher MI compared to all spikes in both MGB and A1 (Figure 8B), indicating 

624 

that enhanced information encoding is not specific to synchronized states under slow 

625 

oscillations. Additionally, a subset of recordings did not show strong slow oscillations in either 

626 

stimulus-driven or spontaneous activity (n(animals): MGB = 4, A1 = 6; n(recordings): MGB = 5, 

627 

A1 = 8). We also examined the stability of cNEs across and within stimulus conditions in these 

628 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

33 

recordings. The majority of cNEs were stable both within and across stimulus conditions (Figure 

629 

8C). In summary, cNEs exist in both MGB and A1 without the presence of strong slow 

630 

oscillations in neural activity. Their enhanced information encoding and stability across stimulus 

631 

conditions are not due to a special behavior of neurons in synchronized states under slow 

632 

oscillations. 

633 

 

634 

Discussion 

635 

This study aimed to investigate whether the auditory thalamus (MGB) contains coordinated 

636 

neural ensembles (cNEs) with enhanced information properties, similar to those observed in A1. 

637 

Our results confirm the presence of cNEs in the MGB, with consistent compositions across 

638 

various, but especially smaller, bin sizes and stable structures across different stimulus 

639 

conditions. Importantly, coordinated spikes among cNE member neurons exhibit higher 

640 

reliability and convey more stimulus-related information than individual neurons.  Neuronal 

641 

groups formed by shared firing-rate changes to stimuli appear not to be congruent with cNEs. 

642 

Furthermore, our findings demonstrate that cNEs are not the result of false positive detections or 

643 

byproducts of slow state oscillations in anesthetized animals. These findings provide support for 

644 

the notion that synchronized neuronal ensembles represent a general principle of local 

645 

organization for information processing in the auditory forebrain. 

646 

cNEs are ubiquitous in local circuit organization 

647 

Neuronal ensembles were proposed as fundamental units for information processing in the brain 

648 

(Hebb, 1949; Buzsáki, 2010), supported by evidence of precise temporal coordination in cortical 

649 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

34 

columns (Atencio and Schreiner, 2013; See et al., 2018; Lankarany et al., 2019). Cortical 

650 

columns consist of neurons with fairly homogeneous properties maintained through intracortical 

651 

processing and shared afferent input (Mountcastle, 1997). This raises the question of whether 

652 

cNEs are unique to cortical organization or represent a general organizational and information 

653 

processing unit along sensory pathways. In the auditory system, reciprocal connectivity exists 

654 

between the MGB and A1, with convergence of frequency tuning and spectral and temporal 

655 

modulation preferences, preserving topographic organization in both regions (Miller et al., 2001, 

656 

2002; Bartlett and Wang, 2007; Read et al., 2011). Therefore, investigating neuronal 

657 

coordination in the MGB, where neurons possess similar properties but differ in their 

658 

organizational and cytoarchitectonic patterns from A1 (Winer, 2010), can provide insights into 

659 

whether cNEs are general organizational principles of local circuits or specialized units specific 

660 

to the cortical circuit composition. 

661 

We conducted recordings of neuronal activity across multiple iso-frequency layers of the MGB 

662 

and were able to reliably detect cNEs in MGB (Figure 2). Neurons within the same cNE 

663 

displayed closer spatial proximity and shared more similar tuning properties (Figure 5D and E), 

664 

indicating functional coherence within cNEs. It is important to note that our recordings were 

665 

limited to relatively small populations of 10-30 neurons due to the techniques employed. 

666 

Therefore, the confinement of spatial and frequency tuning properties within cNEs may vary 

667 

when larger populations with hundreds or thousands of neurons are recorded. 

668 

We further demonstrated that the identification of cNEs relies on the temporal coordination 

669 

among neurons. When the original temporal order among neurons was disrupted through circular 

670 

shuffling of spike trains, a significantly lower number of cNEs were detected in both the MGB 

671 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

35 

and A1 (Figure 4F and 7D). Moreover, the few false positive cNEs that were identified did not 

672 

exhibit the properties observed in cNEs identified in the real data, such as stability across 

673 

different stimulus conditions (Figure 4E and 7E). These findings provide strong support for the 

674 

critical role of temporal coordination in the formation and characterization of cNEs in the 

675 

auditory thalamus and cortex. 

676 

Time scale of cNEs 

677 

Previous studies have investigated neuronal synchrony and coordination across widely differing 

678 

timescales, ranging from a few milliseconds (Lankarany et al., 2019; Shahidi et al., 2019; El-

679 

Gaby et al., 2021) to several hundred milliseconds (Miller et al., 2014; Tremblay et al., 2015; 

680 

Filipchuk et al., 2022). The selection of a specific timescale in these studies was influenced by 

681 

various factors, including the temporal resolution of the recording methods used, the targeted 

682 

functional timescale, and the inter-neuronal distance under investigation. In the context of 

683 

auditory processing, where information changes rapidly within tens of milliseconds (Rosen, 

684 

1992; Lewicki, 2002), we specifically chose a temporal resolution of 10ms. This choice aligns 

685 

with the timescale at which auditory information operates and holds relevance for synaptic 

686 

integration. Selecting an appropriate timescale is crucial for future investigations into the 

687 

functional role of cNEs in synaptic transmission within the auditory thalamocortical system. 

688 

We have demonstrated the robustness of cNE identification across different time bin sizes 

689 

(Figure 3), which can be attributed to the sparse nature of neural activity. Since most 

690 

synchronized neuronal firing occurs at frequencies below 10 Hz (O’Connor et al., 2010), the 

691 

choice of time windows, whether 10ms or 20ms, has minimal impact on the observed 

692 

correlations among neurons. However, it is important to note that cNEs identified with longer 

693 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

36 

time resolutions (hundreds of milliseconds) may significantly differ from those identified with 

694 

shorter resolutions (tens of milliseconds). Specifically, cNEs identified with larger time bins may 

695 

falsely include 'synchronous' events from bursting or rebound activity rather than from an initial 

696 

period that dominates the transmission of stimulus-triggered information. Long synchronization 

697 

windows may also include neurons displaying weaker synchronization within short time 

698 

windows, while potentially de-emphasizing neurons with high temporal precision in synchrony 

699 

(Figure 3D). Hence, the chosen temporal resolution influences the composition and properties of 

700 

cNEs, emphasizing the importance of selecting the appropriate timescale for studying neural 

701 

ensembles. 

702 

Stability of cNEs for spontaneous and evoked activity 

703 

Our study revealed that a significant proportion of cNEs (55% in MGB and 76% in A1) 

704 

maintained a consistent composition during both spontaneous and sensory-evoked neural activity 

705 

(Figure 4E and 7E). This suggests that cNEs generally represent stable configurations within 

706 

local circuits that can manifest independently of stimulus-driven synchrony. These findings align 

707 

with previous research demonstrating similarities between patterns observed in spontaneous and 

708 

stimulus-driven activity (Luczak et al., 2009). Moreover, the similarity between MGB and A1 

709 

cNEs indicates that functional network units are not limited to cortical organization but likely 

710 

exist as a common modality across multiple stages of the sensory pathway. 

711 

cNEs enhance stimulus encoding 

712 

Considering that cNEs were observed in both spontaneous and stimulus-driven activity, some 

713 

argue that they are merely a reflection of background activity and not involved in stimulus 

714 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

37 

encoding and even potentially impairing it (Zohary et al., 1994; Abbott and Dayan, 1999; 

715 

Jermakowicz et al., 2009). Contrary to this notion, our observations revealed that cNE spikes 

716 

exhibit a higher signal-to-noise ratio and convey more information per spike when compared to 

717 

the entire spike train (Figure 6B and C). This suggests that cNE events are more stimulus-

718 

selective than the contributing neurons (See et al., 2021) and exhibit a more reliable response to 

719 

the stimulus features represented by the cNE STRF. The stable connectivity pattern revealed by 

720 

spontaneous, intrinsic activity likely reveals aspects that have been imprinted by extensive 

721 

experience and the behavioral relevance of the associated functional preferences. 

722 

Additionally, we observed an enhanced information encoding in cNE groups and cNE spikes 

723 

when compared to non-cNE groups or coincident spikes, with control of the total number of 

724 

neurons in the groups (Figure 6B and C). This observation suggests that the information increase 

725 

relies on the coordination among cNE member neurons, rather than being a simple result of 

726 

independent population coding (deCharms, 1998; Hatsopoulos et al., 1998).  

727 

Correlated spikes can enhance the transmission of salient auditory information by synchronously 

728 

converging onto their targets (Stevens and Zador, 1998; Zandvakili and Kohn, 2015). 

729 

Additionally, neurons exhibit a multiplexed nature of stimulus encoding, where spikes from the 

730 

same neuron can carry information related to distinct stimulus aspects (Walker et al., 2011; 

731 

Lankarany et al., 2019; See et al., 2021). The function of cNEs may involve selectively choosing 

732 

spikes from member neurons that are most relevant for a specific target information and 

733 

enhancing information propagation while excluding functionally irrelevant spikes of the same 

734 

neurons. This mechanism significantly improves both the robustness and capacity of information 

735 

encoded within a population of neurons (Walker et al., 2011; See et al., 2021). Thus, the 

736 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

38 

presence of cNEs and their coordination within a neuronal population can facilitate efficient 

737 

information processing and transmission in the auditory system. Future studies involving 

738 

simultaneous recordings from two stations along the auditory pathway will be necessary to test 

739 

this hypothesis. 

740 

cNE formation does not depend on strong slow oscillations 

741 

Slow oscillations are commonly observed in neural activity during anesthesia (Chauvette et al., 

742 

2011; Dasilva et al., 2021) and have been shown to influence stimulus encoding (Pachitariu et 

743 

al., 2015). Concerns have been raised regarding whether cNEs are solely a result of anesthesia-

744 

induced synchrony. However, our research findings refute this notion. We focused on a distinct 

745 

time scale of synchronization unrelated to anesthesia-induced slow oscillations and successfully 

746 

detected cNEs in recordings without strong slow oscillations. These cNEs exhibited stable 

747 

structures and enhanced information properties, indicating that they are not solely a byproduct of 

748 

anesthesia-induced synchrony. While we ruled out slow oscillations as the primary force 

749 

underlying cNE formation, it is important to consider their potential interaction with other 

750 

oscillatory activity, such as gamma rhythms (Oberto et al., 2021). Further research is needed to 

751 

explore the interplay between cNEs and different types of brain oscillations. 

752 

 

 

753 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

40 

References 

755 

Abbott LF, Dayan P (1999) The effect of correlated variability on the accuracy of a population 

756 

code. Neural Comput 11:91–101. 

757 

Anderson LA, Linden JF (2011) Physiological differences between histologically defined 

758 

subdivisions in the mouse auditory thalamus. Hear Res 274:48–60. 

759 

Aschauer DF, Eppler JB, Ewig L, Chambers AR, Pokorny C, Kaschube M, Rumpel S (2022) 

760 

Learning-induced biases in the ongoing dynamics of sensory representations predict 

761 

stimulus generalization. Cell Rep 38:110340. 

762 

Atencio CA, Schreiner CE (2008) Spectrotemporal processing differences between auditory 

763 

cortical fast-spiking and regular-spiking neurons. J Neurosci 28:3897–3910. 

764 

Atencio CA, Schreiner CE (2010) Laminar diversity of dynamic sound processing in cat primary 

765 

auditory cortex. J Neurophysiol 103:192–205. 

766 

Atencio CA, Schreiner CE (2013) Auditory cortical local subnetworks are characterized by 

767 

sharply synchronous activity. J Neurosci 33:18503–18514. 

768 

Baeg EH, Kim YB, Huh K, Mook-Jung I, Kim HT, Jung MW (2003) Dynamics of population 

769 

code for working memory in the prefrontal cortex. Neuron 40:177–188. 

770 

Bartlett EL (2013) The organization and physiology of the auditory thalamus and its role in 

771 

processing acoustic features important for speech perception. Brain Lang 126:29–48. 

772 

Bartlett EL, Wang X (2007) Neural representations of temporally modulated signals in the 

773 

auditory thalamus of awake primates. J Neurophysiol 97:1005–1017. 

774 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

41 

Bathellier B, Ushakova L, Rumpel S (2012) Discrete neocortical dynamics predict behavioral 

775 

categorization of sounds. Neuron 76:435–449. 

776 

Bizley JK, Walker KMM, King AJ, Schnupp JWH (2010) Neural ensemble codes for stimulus 

777 

periodicity in auditory cortex. J Neurosci 30:5078–5091. 

778 

Boucly CJ, Pompili MN, Todorova R, Leroux EM, Wiener SI, Zugaro M (2022) Flexible 

779 

communication between cell assemblies and ‘reader’neurons. bioRxiv:2009–2022. 

780 

Brosch M, Schreiner CE (1999) Correlations between neural discharges are related to receptive 

781 

field properties in cat primary auditory cortex. Eur J Neurosci. 11:3517-3530. 

782 

Buzsáki G (2010) Neural Syntax: Cell Assemblies, Synapsembles, and Readers. Neuron 68:362–

783 

385. 

784 

Carrillo-Reid L, Miller JEK, Hamm JP, Jackson J, Yuste R (2015) Endogenous sequential 

785 

cortical activity evoked by visual stimuli. J Neurosci 35:8813–8828. 

786 

Chamberland S, Yang HH, Pan MM, Evans SW, Guan S, Chavarha M, Yang Y, Salesse C, Wu 

787 

H, Wu JC (2017) Fast two-photon imaging of subcellular voltage dynamics in neuronal 

788 

tissue with genetically encoded indicators. Elife 6:e25690. 

789 

Chauvette S, Crochet S, Volgushev M, Timofeev I (2011) Properties of slow oscillation during 

790 

slow-wave sleep and anesthesia in cats. J Neurosci 31:14998–15008. 

791 

Contreras D, Timofeev I, Steriade M (1996) Mechanisms of long-lasting hyperpolarizations 

792 

underlying slow sleep oscillations in cat corticothalamic networks. J Physiol 494:251–264. 

793 

Cowan RL, Wilson CJ (1994) Spontaneous firing patterns and axonal projections of single 

794 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

42 

corticostriatal neurons in the rat medial agranular cortex. J Neurophysiol 71:17–32. 

795 

D’amour JA, Froemke RC (2015) Inhibitory and excitatory spike-timing-dependent plasticity in 

796 

the auditory cortex. Neuron 86:514–528. 

797 

Dan Y, Alonso JM, Usrey WM, Reid RC (1998) Coding of visual information by precisely 

798 

correlated spikes in the lateral geniculate nucleus. Nat Neurosci 1:501–507. 

799 

Dasilva M, Camassa A, Navarro-Guzman A, Pazienti A, Perez-Mendez L, Zamora-López G, 

800 

Mattia M, Sanchez-Vives M V (2021) Modulation of cortical slow oscillations and 

801 

complexity across anesthesia levels. Neuroimage 224:117415. 

802 

deCharms RC (1998) Information coding in the cortex by independent or coordinated 

803 

populations. Proc Natl Acad Sci 95:15166–15168. 

804 

deCharms RC, Merzenich MM (1996) Primary cortical representation of sounds by the 

805 

coordination of action-potential timing. Nature 381:610–613. 

806 

Domanski APF, Kucewicz MT, Russo E, Tricklebank MD, Robinson ESJ, Durstewitz D, Jones 

807 

MW (2023) Distinct hippocampal-prefrontal neural assemblies coordinate memory 

808 

encoding, maintenance, and recall. Curr Biol 33:1220–1236. 

809 

Ebrahimi S, Lecoq J, Rumyantsev O, Tasci T, Zhang Y, Irimia C, Li J, Ganguli S, Schnitzer MJ 

810 

(2022) Emergent reliability in sensory cortical coding and inter-area communication. Nature 

811 

605:713–721. 

812 

Eggermont JJ (2000) Sound-induced synchronization of neural activity between and within three 

813 

auditory cortical areas. (2000) J. Neurophysiol.83:2708-2722. 

814 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

43 

El-Gaby M, Reeve HM, Lopes-dos-Santos V, Campo-Urriza N, Perestenko P V, Morley A, 

815 

Strickland LAM, Lukács IP, Paulsen O, Dupret D (2021) An emergent neural coactivity 

816 

code for dynamic memory. Nat Neurosci 24:694–704. 

817 

Escabí MA, Schreiner CE (2002) Nonlinear Spectrotemporal Sound Analysis by Neurons in the 

818 

Auditory Midbrain. J Neurosci 22:4114–4131. 

819 

Filipchuk A, Schwenkgrub J, Destexhe A, Bathellier B (2022) Awake perception is associated 

820 

with dedicated neuronal assemblies in the cerebral cortex. Nat Neurosci 25:1327–1338. 

821 

Harris KD, Csicsvari J, Hirase H, Dragoi G, Buzsáki G (2003) Organization of cell assemblies in 

822 

the hippocampus. Nature 424:552–556. 

823 

Hasenstaub A, Sachdev RNS, McCormick DA (2007) State changes rapidly modulate cortical 

824 

neuronal responsiveness. J Neurosci 27:9607–9622. 

825 

Hatsopoulos NG, Ojakangas CL, Paninski L, Donoghue JP (1998) Information about movement 

826 

direction obtained from synchronous activity of motor cortical neurons. Proc Natl Acad Sci 

827 

U S A 95:15706–15711. 

828 

Hebb DO (1949) The organization of behavior. Wiley. 

829 

Homma NY, Hullett PW, Atencio CA, Schreiner CE (2020) Auditory Cortical Plasticity 

830 

Dependent on Environmental Noise Statistics. Cell Rep 30:4445-4458.e5. 

831 

Ince RAA, Panzeri S, Kayser C (2013) Neural codes formed by small and temporally precise 

832 

populations in auditory cortex. J Neurosci 33:18277–18287. 

833 

Jermakowicz WJ, Chen X, Khaytin I, Bonds AB, Casagrande VA (2009) Relationship between 

834 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

44 

spontaneous and evoked spike-time correlations in primate visual cortex. J Neurophysiol 

835 

101:2279–2289. 

836 

Konorski J (1948) Conditioned reflexes and neuron organization. 

837 

Kreiter AK, Singer W (1996) Stimulus-dependent synchronization of neuronal responses in the 

838 

visual cortex of the awake macaque monkey. J Neurosci 16:2381–2396. 

839 

Lankarany M, Al-Basha D, Ratté S, Prescott SA (2019) Differentially synchronized spiking 

840 

enables multiplexed neural coding. Proc Natl Acad Sci U S A 116:10097–10102. 

841 

Laubach M, Wessberg J, Nicolelis MAL (2000) Cortical ensemble activity increasingly predicts 

842 

behaviour outcomes during learning of a motor task. Nature 405:567–571. 

843 

Léger JF, Stern EA, Aertsen A, Heck D (2005) Synaptic integration in rat frontal cortex shaped 

844 

by network activity. J Neurophysiol 93:281–293. 

845 

Lopes-dos-Santos V, Ribeiro S, Tort ABL (2013) Detecting cell assemblies in large neuronal 

846 

populations. J Neurosci Methods 220:149–166. 

847 

Luczak A, Barthó P, Harris KD (2009) Spontaneous Events Outline the Realm of Possible 

848 

Sensory Responses in Neocortical Populations. Neuron 62:413–425. 

849 

Marchenko VA, Pastur LA (1967) Distribution of eigenvalues for some sets of random matrices. 

850 

Mat Sb 114:507–536. 

851 

Merzenich MM, Knight PL, Roth GL, Linn G (2013) Representation of cochlea within primary 

852 

auditory cortex in the cat of Cochlea in the Cat. :231–249. 

853 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

45 

Metherate R, Ashe JH (1993) Nucleus basalis stimulation facilitates thalamocortical synaptic 

854 

transmission in the rat auditory cortex. Synapse 14:132–143. 

855 

Miller JEK, Ayzenshtat I, Carrillo-Reid L, Yuste R (2014) Visual stimuli recruit intrinsically 

856 

generated cortical ensembles. Proc Natl Acad Sci U S A 111:E4053–E4061. 

857 

Miller LM, Escabí MA, Read HL, Schreiner CE (2001) Functional convergence of response 

858 

properties in the auditory thalamocortical system. Neuron 32:151–160. 

859 

Miller LM, Escabí MA, Read HL, Schreiner CE (2002) Spectrotemporal receptive fields in the 

860 

lemniscal auditory thalamus and cortex. J Neurophysiol 87:516–527. 

861 

Miller LM, Recanzone GH (2009) Populations of auditory cortical neurons can accurately 

862 

encode acoustic space across stimulus intensity. Proc Natl Acad Sci 106:5931–5935. 

863 

Miller LM, Schreiner CE (2000) Stimulus-based state control in the thalamocortical system. J 

864 

Neurosci 20:7011–7016. 

865 

Mochol G, Hermoso-Mendizabal A, Sakata S, Harris KD, De La Rocha J (2015) Stochastic 

866 

transitions into silence cause noise correlations in cortical circuits. Proc Natl Acad Sci U S 

867 

A 112:3529–3534. 

868 

Mogensen H, Norrlid J, Enander JMD, Wahlbom A, Jörntell H (2019) Absence of repetitive 

869 

correlation patterns between pairs of adjacent neocortical neurons in vivo. Front Neural 

870 

Circuits 13:1–11. 

871 

Morel A, Rouiller E, de Ribaupierre Y, de Ribaupierre F (1987) Tonotopic organization in the 

872 

medial geniculate body (MGB) of lightly anesthetized cats. Exp Brain Res 69:24–42. 

873 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

46 

Mountcastle VB (1997) The columnar organization of the neocortex. Brain a J Neurol 120:701–

874 

722. 

875 

Neske GT (2016) The slow oscillation in cortical and thalamic networks: Mechanisms and 

876 

functions. Front Neural Circuits 9:1–25. 

877 

Niessing J, Friedrich RW (2010) Olfactory pattern classification by discrete neuronal network 

878 

states. Nature 465:47–52. 

879 

O’Connor DH, Peron SP, Huber D, Svoboda K (2010) Neural activity in barrel cortex underlying 

880 

vibrissa-based object localization in mice. Neuron 67:1048–1061. 

881 

Oberto VJ, Boucly CJ, Gao H, Todorova R, Zugaro MB, Wiener SI (2022) Distributed cell 

882 

assemblies spanning prefrontal cortex and striatum. Curr Biol:1–13. 

883 

Pachitariu M, Lyamzin DR, Sahani M, Lesica NA (2015) State-dependent population coding in 

884 

primary auditory cortex. J Neurosci 35:2058–2073. 

885 

Pachitariu M, Sridhar S, Stringer C (2023) Solving the spike sorting problem with Kilosort. 

886 

bioRxiv:2023.01.07.523036. 

887 

Paninski L, Shoham S, Fellows MR, Hatsopoulos NG, Donoghue JP (2004) Superlinear 

888 

population encoding of dynamic hand trajectory in primary motor cortex. J Neurosci 

889 

24:8551–8561. 

890 

Peyrache A, Benchenane K, Khamassi M, Wiener SI, Battaglia FP (2010) Principal component 

891 

analysis of ensemble recordings reveals cell assemblies at high temporal resolution. J 

892 

Comput Neurosci 29:309–325. 

893 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

47 

Polley DB, Read HL, Storace DA, Merzenich MM (2007) Multiparametric auditory receptive 

894 

field organization across five cortical fields in the albino rat. J Neurophysiol 97:3621–3638. 

895 

Qiu A, Schreiner CE, Escabí MA (2003) Gabor analysis of auditory midbrain receptive fields: 

896 

Spectro-temporal and binaural composition. J Neurophysiol 90:456–476. 

897 

Read HL, Nauen DW, Escabí MA, Miller LM, Schreiner CE, Winer JA (2011) Distinct core 

898 

thalamocortical pathways to central and dorsal primary auditory cortex. Hear Res 274:95–

899 

104. 

900 

Rosen S (1992) Temporal information in speech: acoustic, auditory and linguistic aspects. Philos 

901 

Trans R Soc London Ser B Biol Sci 336:367–373. 

902 

Sanchez-Vives MV, McCormick DA (2000) Cellular and network mechanisms of rhythmic 

903 

recurrent activity in neocortex. Nat Neurosci 3:1027–1034. 

904 

See JZ, Atencio CA, Sohal VS, Schreiner CE (2018) Coordinated neuronal ensembles in primary 

905 

auditory cortical columns. Elife 7:1–33. 

906 

See JZ, Homma NY, Atencio CA, Sohal VS, Schreiner CE (2021) Information diversity in 

907 

individual auditory cortical neurons is associated with functionally distinct coordinated 

908 

neuronal ensembles. Sci Rep 11:1–15. 

909 

Shahidi N, Andrei AR, Hu M, Dragoi V (2019) High-order coordination of cortical spiking 

910 

activity modulates perceptual accuracy. Nat Neurosci 22:1148–1158. 

911 

Smith PH, Uhlrich DJ, Manning KA, Banks MI (2012) Thalamocortical projections to rat 

912 

auditory cortex from the ventral and dorsal divisions of the medial geniculate nucleus. J 

913 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

48 

Comp Neurol 520:34–51. 

914 

Steinmetz NA et al. (2021) Neuropixels 2.0: A miniaturized high-density probe for stable, long-

915 

term brain recordings. Science (80- ) 372. 

916 

Steriade M, McCormick DA, Sejnowski TJ (1993) Thalamocortical oscillations in the sleeping 

917 

and aroused brain. Science (80- ) 262:679–685. 

918 

Stevens CF, Zador AM (1998) Input synchrony and the irregular firing of cortical neurons. Nat 

919 

Neurosci 1:210–217. 

920 

Theunissen FE, Sen K, Doupe AJ (2000) Spectral-temporal receptive fields of nonlinear auditory 

921 

neurons obtained using natural sounds. J Neurosci 20:2315–2331. 

922 

Tremblay S, Pieper F, Sachs A, Martinez-Trujillo J (2015) Attentional filtering of visual 

923 

information by neuronal ensembles in the primate lateral prefrontal cortex. Neuron 85:202–

924 

215. 

925 

Wahlbom A, Mogensen H, Jörntell H (2021) Widely Different Correlation Patterns Between 

926 

Pairs of Adjacent Thalamic Neurons In vivo. Front Neural Circuits 15:1–10. 

927 

Walker KMM, Bizley JK, King AJ, Schnupp JWH (2011) Multiplexed and robust 

928 

representations of sound features in auditory cortex. J Neurosci 31:14565–14576. 

929 

Wills TJ, Lever C, Cacucci F, Burgess N, O’Keefe J (2005) Attractor dynamics in the 

930 

hippocampal representation of the local environment. Science (80- ) 308:873–876. 

931 

Winer JA (2010) Neurochemical organization of the medial geniculate body and auditory cortex. 

932 

In: The auditory cortex, pp 209–234. Springer. 

933 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

49 

Winer JA, Miller LM, Lee CC, Schreiner CE (2005) Auditory thalamocortical transformation: 

934 

Structure and function. Trends Neurosci 28:255–263. 

935 

Yoshida T, Ohki K (2020) Natural images are reliably represented by sparse and variable 

936 

populations of neurons in visual cortex. Nat Commun 11:872. 

937 

Yuste R (2015) From the neuron doctrine to neural networks. Nat Rev Neurosci 16:487–497. 

938 

Zandvakili A, Kohn A (2015) Coordinated Neuronal Activity Enhances Corticocortical 

939 

Communication. Neuron 87:827–839. 

940 

Zohary E, Shadlen MN, Newsome WT (1994) Correlated neuronal discharge rate and its 

941 

implications for psychophysical performance. Nature 370:140–143. 

942 

Figure Legends 

943 

Figure 1. In vivo recordings in rat MGB 

944 

(A) Left: Schematic of the recording setup in the MGB using a linear 64-channel probe. (i) and 

945 

(ii): two electrode penetrations with multi-unit (MU) recordings from the MGB. Left: Stacked 

946 

firing rate (color coded) of pure-tone frequency response areas. Right: Characteristic frequencies 

947 

(CF, the frequency at which the response threshold is the lowest). The red dashed lines indicate 

948 

the potential boundaries of the ventral MGB. (B) Example STRFs of single units (SUs) from a 

949 

recording in the MGB. Unit numbers 1 to 3 indicate the positions and STRFs of pairs of neurons 

950 

whose CCGs are plotted in (C). (C) Example CCGs from two pairs of neurons (#1 - #3 and #2 - 

951 

#3). The black bars represent the CCGs of stimulus-driven activity, while the grey lines represent 

952 

the CCGs of spontaneous activity. The baseline is estimated by averaging the counts in 5ms 

953 

windows at the shoulders of the CCGs and is indicated by dashed red lines. 

954 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

50 

Figure 2 Groups of neurons with coordinated activities exist in MGB 

955 

(A) Procedures for detecting cNEs in a thalamic penetration. (i) Correlation matrix of spike 

956 

trains. (ii) Eigenvalues of the correlation matrix shown in (i). The dashed red line represents the 

957 

99.5th percentile of the Marchenko–Pastur distribution, which was used as the significance 

958 

threshold for eigenvalues. The top four eigenvalues are significant and represent the number of 

959 

detected cNEs. (iii) IC weights of neurons for each cNE. The green dots represent neurons that 

960 

are members of a cNE. (iv) cNE members (red stems) are neurons with IC weights exceeding the 

961 

threshold (1

/

N

) shown as grey areas. (v) Example of cNE activation. Top: Activity trace of cNE 

962 

#1. The red line shows the threshold estimated using Monte Carlo methods. The peaks crossing 

963 

the threshold indicate cNE events when multiple cNE member neurons fire jointly. Bottom: 

964 

Spike raster of neurons, with red ticks indicating spikes that contribute to instances of cNE 

965 

events, which were referred to as cNE spikes. Shaded areas show member neurons. (B) 

966 

Correlations (10ms bin) of neuron pairs that were both members of the same cNE (members) or 

967 

neuron pairs that were not members of the same cNE (non-members) in MGB (p = 6.9e-235, 

968 

Mann–Whitney U test). (C) Z-scored CCGs (1ms bin) of member pairs (left) and non-member 

969 

pairs (right) in MGB. Top: Stacked CCGs ordered by the peak delay. Bottom: Average of the 

970 

data above (mean ± SD; shaded area: p < 0.01, permutation test, shuffling the members/non-

971 

members labels).  

972 

Figure 3. Variability of IC weights across different time bin sizes. 

973 

(A) Example of two MGB cNEs whose member neurons are either consistently identified across 

974 

different bin sizes (i) or only detected using smaller bin sizes (ii). (B) Proportion of significantly 

975 

matched cNEs. Using different bin sizes as reference bin sizes (row), we calculated the 

976 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

51 

proportion of cNEs on different bin sizes that have a significantly matched cNE on the reference 

977 

bin size. The red square highlights the proportion of 10ms cNEs that have significant matches 

978 

with 160ms cNEs and is further analyzed in (C). (C) Left: Correlation of IC weights of 10ms 

979 

cNEs with the most correlated IC weights of 160ms cNEs. Right: Proportion of shared 

980 

membership between 10ms cNEs with their most correlated 160ms cNEs. (D) Firing rate of 

981 

member neurons in 10ms-only cNEs, 10ms cNEs without a significant match with 160ms cNEs, 

982 

and 160ms-only cNEs (p = 0.57, Kolmogorov-Smirnov test). (E) Top: Schematic of membership 

983 

of neurons for different bin sizes. Bottom: Mean z-scored CCGs (1ms bin, mean ± SD) between 

984 

(i) member neurons in both 10ms and 160ms cNEs, (ii) member neurons only in the 160ms cNE 

985 

and member neurons in both 10ms and 160ms cNEs (shaded area: p < 0.01, permutation test, 

986 

shuffling neuron pair labels of i and ii), and (iii) member neurons only in the 10ms cNE and 

987 

member neurons in both 10ms and 160ms cNEs (no time bin showed significant difference from 

988 

i). 

989 

Figure  4.  cNEs  identified  in  spontaneous  activity  are  mostly  preserved  in  stimulus-driven 

990 

activity. 

991 

(A) Diagram illustrating the recording sequence and partitioning of spontaneous (yellow/orange) 

992 

and DMR-evoked (dark green/light green) activity. The four blocks allowed the comparison of 

993 

cNEs obtained within stimulus conditions and across stimulus conditions. (B) Absolute 

994 

correlation values of the IC weights calculated on adjacent blocks from an MGB recording 

995 

including the two examples (i and ii) shown in (C). (C) Example IC weights on adjacent blocks 

996 

with high (i) and moderate (ii) correlation values across stimulus conditions. The dashed lines 

997 

show the threshold to determine the membership of the neurons. (D) The two cNE examples in 

998 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

52 

(C) have significantly matched IC weights across stimulus conditions. See method for how the 

999 

null distribution was generated. The significance threshold for the correlation values was set at p 

1000 

= 0.01 (red dashed line, 99.5th percentile of the null distribution). The brown (i) and blue (ii) 

1001 

solid lines represent the two examples in (C). (E) Correlations of IC weights identified on 

1002 

adjacent activity blocks for real (red) and circularly shifted data (pink). The hollow histograms 

1003 

show all correlations of matched cNEs on adjacent blocks; the histograms show significant 

1004 

correlations based on the test shown in (D). The inset numbers show the percentage of cNEs with 

1005 

significantly matched IC weights on adjacent blocks. The triangles show the median of all IC 

1006 

weight correlations (real data vs shuffled data: spon, p = 5.6e-36; dmr: p = 1.3e-28; cross, p = 

1007 

2.0e-22, Mann–Whitney U test with Bonferroni correction). (F) Number of cNEs detected using 

1008 

real and circularly shifted activities on the four recording blocks (spon1, p = 4.7e-10; spon2, p = 

1009 

4.7e-10; dmr1, p = 4.7e-10; dmr2, p = 4.7e-10, Wilcoxon signed-rank test with Bonferroni 

1010 

correction). 

1011 

Figure 5. Properties of cNEs in MGB. 

1012 

(A) Number of cNEs detected in any given penetration increases with the number of recorded 

1013 

neurons. (B) cNE size increases with the number of recorded neurons. (C) The number of cNEs a 

1014 

neuron belongs to. (D) Spatial distribution of cNE members. (i) Pairwise distance of neurons in 

1015 

the same cNE (colored bar) or neurons not in the same cNE (black line). (ii) Spatial span of cNE 

1016 

members (colored bar) and random groups of neurons with the same number of neurons as cNEs 

1017 

(black line). (E) Frequency tuning distribution of cNE members. (i) Pairwise difference in the 

1018 

best frequencies (BF) of neurons. (ii) The largest difference in the BF among cNE members or 

1019 

random groups of neurons. (Mann–Whitney U test). 

1020 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

53 

Figure 6. MGB cNEs can refine sound features encoded by member neurons. 

1021 

(A) Two example STRFs of MGB neurons calculated with all spikes (left) and cNE spikes 

1022 

(right). All spikes are subsampled to have an equal number as cNE spikes. (B) (i) STRF peak-

1023 

trough difference (PTD) for cNE spikes or all spikes of neurons. (ii) STRF PTD for groups of 

1024 

cNE members or non-members. (iii) STRF PTD for cNE spikes and coincident spikes of a 

1025 

neuron. The coincident spikes refer to instances where a neuron's firing occurs within a 10ms 

1026 

timeframe of another neuron's firing in a group of non-member neurons. This group is designed 

1027 

to match the number of neurons present in the cNE. (C) (i) Mutual information (MI) between 

1028 

stimulus and cNE spikes or all spikes of neurons. (ii) STRF MI for groups of cNE members or 

1029 

non-members. (iii) STRF MI for cNE spikes and coincident spikes of a neuron. (Wilcoxon 

1030 

signed-rank test). 

1031 

Figure 7. MGB and A1 cNEs have similar properties. 

1032 

(A) Left: schematic of the recording setup in A1 using a 2-shank probe with 64 channels. Right: 

1033 

MU responses to pure tones as in Figure 1A. (B) (i) PTD of STRFs calculated using all spikes or 

1034 

only cNE spikes from a neuron in A1. (ii) Difference between cNE spike STRF PTD and all 

1035 

spike STRF PTD in MGB and A1 (p = 0.72, Mann–Whitney U test). (C) (i) MI of STRFs 

1036 

calculated using all spikes or only cNE spikes from a neuron in A1. (ii) Difference between cNE 

1037 

spike STRF MI and all spike STRF MI in MGB and A1 (p = 0.92, Mann–Whitney U test). (D) 

1038 

Number of cNEs detected using real and circularly shifted activities on the four recording blocks 

1039 

in A1, as shown in Figure 4F for MGB. (spon1: p = 6.1e-5, spon2: p = 6.1e-5, dmr1: p = 6.1e-5, 

1040 

dmr2: p = 6.1e-5, n = 17 recordings, Wilcoxon signed-rank test with Bonferroni correction). (E) 

1041 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

54 

Correlations of IC weights identified on adjacent activity blocks for real (blue) and circularly 

1042 

shifted data (light blue) in A1. The hollow histograms and histograms show the distribution of 

1043 

correlations and significant correlations as in Figure 4E. The triangles show the median of IC 

1044 

weight correlations (spon: p = 1.4e-33, dmr: p = 4.6e-34, cross: p = 4.8e-28, Mann–Whitney U 

1045 

test with Bonferroni correction). 

1046 

Figure 8. cNE events do not rely on slow oscillation in neural activity. 

1047 

(A) Silence density and firing rate (FR) coefficient of variation (CV) of population activity in 

1048 

response to DMR. Some recordings show strong slow oscillation in population activity with 

1049 

pronounced silent period between highly active moments (i), while others show little (ii) or 

1050 

moderate (iii) levels of slow MU firing rate oscillation. Recordings with silence density 

0.4 

1051 

and MU firing rate coefficient of variance 

0.8 (dashed lines) did not show prominent slow 

1052 

oscillation in population activity and were included in (B). (B) STRF MI with all spikes and cNE 

1053 

spikes in recordings without prominent slow oscillations (Wilcoxon signed-rank test). (C) 

1054 

Correlation values of cNE IC weights on adjacent activity blocks from recordings with no 

1055 

prominent slow oscillations in both spontaneous and stimulus-driven activities. The inset 

1056 

numbers show the percentage of cNEs with significantly matched IC weights on adjacent blocks 

1057 

in MGB (red) and A1 (blue). The hollow histograms and histograms show all correlation values 

1058 

and significant correlation values with the same presentation scheme as in Figure 4E. 

1059 

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

JNeurosci Accepted Manuscript

JNEUROSCI.1729-23.2024.full-html.html
background image

JNeurosci Accepted Manuscript