Sie sind auf Seite 1von 21

Article

The Locus Coeruleus Is a Complex and


Differentiated Neuromodulatory System
Highlights Authors
d The LC, the global source of forebrain noradrenaline, uses an Nelson K. Totah, Ricardo M. Neves,
ensemble code Stefano Panzeri, Nikos K. Logothetis,
Oxana Eschenko
d Ensembles are active over many timescales including
periodic infra-slow activity Correspondence
d Ensembles consist of neurons spatially distributed nelson.totah@tuebingen.mpg.de (N.K.T.),
oxana.eschenko@tuebingen.
throughout the nucleus
mpg.de (O.E.)
d Local LC interactions and LC-cortical interactions depend on
LC waveform shape
In Brief
Totah et al. record spiking from
populations of noradrenaline neurons in
the brainstem nucleus, locus coeruleus,
which affects brain-wide functions.
Instead of en masse activity typical of
global neuromodulation, these
recordings reveal spatially and
temporally diverse ensembles providing
targeted neuromodulation.

Totah et al., 2018, Neuron 99, 1–14


September 5, 2018 ª 2018 Elsevier Inc.
https://doi.org/10.1016/j.neuron.2018.07.037
Please cite this article in press as: Totah et al., The Locus Coeruleus Is a Complex and Differentiated Neuromodulatory System, Neuron (2018), https://
doi.org/10.1016/j.neuron.2018.07.037

Neuron

Article

The Locus Coeruleus Is a Complex


and Differentiated Neuromodulatory System
Nelson K. Totah,1,* Ricardo M. Neves,1 Stefano Panzeri,2 Nikos K. Logothetis,1,3 and Oxana Eschenko1,4,*
1Department of Physiology of Cognitive Processes, Max Planck Institute for Biological Cybernetics, Max-Planck-Ring 8, 72076 Tuebingen,

Germany
2Laboratory of Neural Computation, Istituto Italiano di Tecnologia, Corso Bettini 31, 38068 Rovereto, Italy
3Division of Imaging Science and Biomedical Engineering, University of Manchester, M13 9PT Manchester, UK
4Lead Contact

*Correspondence: nelson.totah@tuebingen.mpg.de (N.K.T.), oxana.eschenko@tuebingen.mpg.de (O.E.)


https://doi.org/10.1016/j.neuron.2018.07.037

SUMMARY instead, suggested a one-to-one and/or ensemble-based orga-


nization, by showing that target-selective pathways originating
Diffuse projections of locus coeruleus (LC) neurons from discrete LC sub-populations affect different behaviors
and evidence of synchronous spiking have long (Chandler et al., 2014; Hirschberg et al., 2017; Kebschull et al.,
been perceived as features of global neuromodula- 2016; Uematsu et al., 2017; Wagatsuma et al., 2018). These
tion. Recent studies demonstrated the possibility of new findings encourage further investigation of the role of LC
targeted modulation by subsets of LC neurons. Non- sub-population projection specificity and the possibility that
activity within those sub-populations is correlated.
global neuromodulation depends on target specificity
Even with target-selective projections, however, the LC
and the differentiated spatiotemporal dynamics
would only be able to affect different brain regions when its
within LC. Here, we characterized interactions be- neurons are not all synchronized and if they can fire in distinct,
tween 3,164 LC cell pairs in the rat LC under urethane ensemble-based patterns. Previous seminal studies have
anesthesia. Spike count correlations were near zero begun to address synchronous activity in the LC but provide
and only a small proportion of unit pairs had synchro- a partly contrasting and still incomplete view especially with re-
nized spontaneous (15%) or evoked (16%) discharge. gard to beyond-pairwise single-unit ensembles and the spatial
We identified infra-slow (0.01–1 Hz) fluctuations of LC and temporal scales over which synchrony is expressed. Extra-
unit spike rate, which were also asynchronous across cellular recordings of LC spiking in awake and anesthetized rats
the population. Despite overall sparse population and slice recordings of local field and membrane potentials
synchrony, we report the existence of LC ensembles have provided indirect evidence for widespread synchrony (Al-
varez et al., 2002; Aston-Jones and Bloom, 1981a, 1981b; Chen
and relate them to forebrain projection targets. We
and Sara, 2007; Finlayson and Marshall, 1988; Ishimatsu and
also show that spike waveform width was related to
Williams, 1996). For example, sensory stimuli elicit single-unit
ensemble membership, propensity for synchroniza- spikes that are highly synchronous with multi-unit spiking of
tion, and interactions with cortex. Our findings sug- surrounding population (Chen and Sara, 2007). Moreover, the
gest a partly differentiated and target-specific norad- local field potential (LFP, a marker of transmembrane currents
renergic signal. and other peri-synaptic activity), with which multi-unit activity
(MUA) is synchronized, is coherent between spatially segre-
gated sites in the LC and suggests synchronous activity
INTRODUCTION throughout the nucleus (Aston-Jones and Bloom, 1981b,
1981a; Ishimatsu and Williams, 1996). Direct evidence for syn-
Neuromodulation has long been seen as a ‘‘one-to-many’’ activ- chrony was demonstrated using cross-correlogram measures
ity due to diffuse efferents of neuromodulatory nuclei and volume from 23 pairs of single units recorded in monkeys during various
transmission (Mitchell et al., 1994; Mundorf et al., 2001; Room states of vigilance (Usher et al., 1999). However, in vitro studies
et al., 1981; Schwarz et al., 2015). Neuromodulatory nuclei, as suggest that synchrony may be quite low or only increase in the
hubs of the brainstem arousal systems, are often considered presence of specific neurotransmitter (e.g., hypocretin) drive
to be undifferentiated ‘‘state controllers’’ (Aston-Jones and (Christie et al., 1989; Rancic et al., 2018; van den Pol et al.,
Cohen, 2005; Dayan and Yu, 2006; Lee and Dan, 2012). Example 2002). One potential gap in the aforementioned evidence is
par excellence is the noradrenergic locus coeruleus (LC); its that it has relied on indiscriminate multi-unit spiking or record-
diffuse projections are conserved across vertebrates (Aston- ings of only two single units simultaneously. Evidently, such
Jones and Bloom, 1981a; Smeets and González, 2000) and measurements do not allow an assessment of potential
permit regulation of broad forebrain networks related to a multi- ensemble-based spike patterns. Interpreting the emerging per-
tude of functions (Sara and Bouret, 2012). However, recent spectives for targeted neuromodulation by an LC ensemble
studies have challenged this one-to-many configuration and, code would also benefit from knowledge about the spatial

Neuron 99, 1–14, September 5, 2018 ª 2018 Elsevier Inc. 1


Please cite this article in press as: Totah et al., The Locus Coeruleus Is a Complex and Differentiated Neuromodulatory System, Neuron (2018), https://
doi.org/10.1016/j.neuron.2018.07.037

Table 1. The Number of LC Single Units Recorded in Each Rat Extracellular waveforms separated into two groups based on
spike width and after-hyperpolarization amplitude (Figures 1A
Number of Number (%) of Number (%) of
Rat Identifier Single Units Narrow Units Wide Units and 1B). We refer to these populations as ‘‘narrow’’ and
‘‘wide’’ units. Out of 234 units, 34 were narrow (15%) and 200
959.1 7 5 (71) 2 (29)
were wide (85%). Waveform width was significantly different
980.4 9 9 (100) 0 (0)
between narrow and wide units (0.462 ± 0.021 ms versus
105.1 5 0 (0) 5 (100) 1.082 ± 0.010 ms; t test, T(232) = 24.56, p < 0.0001). Waveform
105.2 52 10 (19) 42 (81) amplitude was similar (t test, T(232) = 0.599, p = 0.550), with
109.1 21 0 (0) 21 (100) average amplitudes of 0.136 ± 0.019 mV (narrow) and 0.129 ±
134.2 15 0 (0) 15 (100) 0.004 mV (wide). Both narrow and wide units had identical
259.1 18 0 (0) 18 (100) response profiles to foot shocks with short-latency excitation
259.2 27 4 (15) 23 (85) followed by prolonged inhibition and both were inhibited by
clonidine (Figure S1). Stimulation of forebrain sites elicited
263.2 36 2 (6) 34 (94)
antidromic responses in similar percentages of narrow (30%)
286.3 8 3 (38) 5 (62)
and wide (38%) units (two-sided, Fisher’s exact test, odds ratio =
288.1 6 0 (0) 6 (100) 1.41, CI = [0.349 5.714], p = 0.744). Thus, both subsets of units
288.2 30 1 (3.3) 29 (96.7) can be classified as putative NE-producing, forebrain-projecting
The numbers and percent of each unit separated by waveform width. LC neurons.
Interestingly, waveform width overlapped with a number of
characteristic differences. Narrow units discharged at signifi-
and temporal structure of population activity within the LC and cantly higher rates (Figures 1C and 1D; median and SD, 1.28 ±
its relation to projection specificity of individual LC neurons. In 0.73 spikes/s and 0.64 ± 0.63 spikes/s, respectively; Wilcoxon-
earlier studies, the available tools were technically prohibitive Mann-Whitney, Z = 4.23, p = 0.00002, Cohen’s D = 0.823,
for ensemble recordings because the small size of the LC power = 0.973). Also, narrow units were more predominant in
permitted mainly single-electrode implantations. Additionally, the ventral aspect of the nucleus (Figure 1E). This distribution
the single-unit waveforms of the densely packed LC cell bodies was present in all 12 rats.
are difficult to isolate. Here, we overcame these obstacles by
using a high-density recording array to collect the first large- LC Single Units Have Near Zero Spontaneous and
scale population data from the LC. We recorded, in total, 234 Evoked Spike Count Correlations
units (up to 52 single units per rat) under urethane anesthesia A seminal early study clearly demonstrated highly synchronous
and analyzed 3,164 unit pairs. Stable long-duration recording spiking of LC neurons (80% out of 23 single-unit pairs) in
of single units permitted statistically robust evaluation of spike monkeys during periods of accurate performance of a cognitive
train synchrony. Acute experiments also allowed characteriza- choice task (Usher et al., 1999). Based on this result, pairwise
tion of units’ projection patterns by their antidromic responses spike count correlation coefficients and the proportion of signif-
to stimulation of 15 forebrain sites. Here, we report that the icant cross-correlograms for the LC population are expected to
degree of synchrony in large populations of simultaneously re- greatly exceed the values reported for cortex, which are, respec-
corded LC single units is rather low during a brain state with tively, 0.1 and 32 (16%) of 200 pairs (Cohen and Kohn, 2011;
predominant slow-wave dynamics. We also demonstrate the Engel et al., 1990). In order to estimate correlated firing in the
presence of simultaneously active, yet distinct LC ensembles LC, we calculated spike count correlation coefficients (RSC) for
that exhibit synchrony among neurons over slow (seconds) to 3,164 single-unit pairs in bins of 200 ms and 1 s, which were
fast (sub-millisecond) timescales. Ensemble firing patterns in chosen based on the 100–200 ms duration of coincidental
the LC may allow circuit-based neuromodulation. spiking reported in LC cross-correlograms (Usher et al., 1999)
and the previously demonstrated relationship between rhythmic
RESULTS increases of LC MUA in relation to cortical 1–2 Hz slow waves
(Eschenko et al., 2012; Safaai et al., 2015). RSC values were
LC Single Units Separable by Waveform Width Differ in distributed around zero (Figure 2). The SEM was 0.044 ± 0.001
Spike Rate and Spatial Distribution for 200 ms bins and 0.098 ± 0.003 for 1 s bins. RSC did not
We isolated 234 single units in 12 rats (5–52 units per rat; Table 1). depend on a distance between the units (Figure S2A).
All single units exhibited typical characteristics of putative One factor to consider is that anesthesia may have inflated the
noradrenergic LC cells (Figure S1). Extracellular recordings do measured RSC values, which are higher in the cortex under
not allow unambiguous identification of cell neurochemical iden- various types of anesthesia in both rodents and monkeys as
tity or location. Therefore, we used an agonist (clonidine) for the compared to awake states (Aasebø et al., 2017; Ecker et al.,
alpha-2 adrenergic receptor to demonstrate a complete cessa- 2014; Erchova et al., 2002; Greenberg et al., 2008; Sakata and
tion of firing of the recorded units (Figure S10). Nearby structures Harris, 2009). The increased cortical synchrony is due to tempo-
could have contributed low-amplitude spikes, but they are not ral compacting of cortical population spiking into particular
inhibited by clonidine due to the lack alpha-2 adrenoreceptors periods (up states) of the slow EEG/LFP oscillations that occur
(McCune et al., 1993). We also histologically verified the location during anesthesia or natural sleep. The LC exhibits burst-
of the electrode tract within the LC core (Figure S9). like population firing during cortical slow oscillation states

2 Neuron 99, 1–14, September 5, 2018


Please cite this article in press as: Totah et al., The Locus Coeruleus Is a Complex and Differentiated Neuromodulatory System, Neuron (2018), https://
doi.org/10.1016/j.neuron.2018.07.037

Figure 1. Distinct Populations of LC Single


Units Were Separable by Waveform Width
and Differed in Spike Rate and Spatial
Distribution
(A) The average waveforms of five example units
illustrate the diversity of waveform shapes.
(B) Units were separable based on the waveform
duration and the amplitude of the after-hyperpo-
larization in relation to the first peak. The green and
blue asterisks refer to the example waveforms in
(A) with the same markings.
(C and D) Scatterplots with the mean spontaneous
spike rate and inter-spike interval for each LC unit.
The insets show the mean and SE for each subset
of units. n = 34 (narrow) and 200 (wide) units.
(E) Narrow units were predominantly distributed in
the ventral aspect of the LC and sparsely recorded
elsewhere in the nucleus. The mean and SE of the
probability distribution are plotted.
See also Figures S1 and S10.

Sensory stimuli evoke burst spiking of


LC neurons, which may be a homoge-
neous population response to each stim-
ulus (Aston-Jones and Bloom, 1981b). To
study correlations between evoked LC
spiking activity, we applied a single foot
shock (5.0 mA, 0.5 ms pulse duration)
and measured spike count correlations
(Aston-Jones and Bloom, 1981a; Eschenko et al., 2012; Lesti- during the time window of the maximal evoked discharge
enne et al., 1997; Safaai et al., 2015) and may therefore also (50 ms after stimulation; see red spikes in Figure 3A). We used
have higher inter-neuronal synchrony during the states with the a stimulation intensity that has been shown to elicit a population
prevalence of cortical slow LFP oscillations. In line with this response (Cedarbaum and Aghajanian, 1978; Devilbiss and
prediction, we found that LC pairwise correlations were higher Berridge, 2006; Mana and Grace, 1997; Neves et al., 2018;
when the cortex was in a more synchronized (LFP slow oscilla- Uematsu et al., 2017; Valentino and Foote, 1987). Mean evoked
tion) compared to a less synchronized (‘‘activated’’) state (Fig- RSC values were distributed around zero (Figure 2). Increasing
ure S3). To account for the role of network activity, we developed the stimulus intensity to 5 pulses (at 30 Hz) did not increase syn-
a non-parametric permutation test of significance that discounts chrony (single pulse, 0.007 ± 0.006 versus 30 Hz, 0.012 ± 0.006,
spurious correlations due to non-random inter-spike intervals Wilcoxon-Mann-Whitney, Z = 1.24, p = 0.214). Moreover,
and common locking to, for example, cortical slow oscillations. accounting for possible adaptation by calculating correlations
Only 16% of pairs had synchronous spiking that was significantly in blocks of 5 trials (e.g., trials 1–5, 6–10, and so on) did not
higher than what would be expected to occur by chance (one- demonstrate any propensity for higher correlations during earlier
sided permutation test, p < 0.01). stimulation trials for either stimulation intensity (single pulse,
Another important factor to consider is anesthesia-induced Kruskal-Wallis H = 0.68, p = 0.711; 30 Hz, Kruskal-Wallis
suppression of firing rate because lower firing rates lead to lower H = 0.69, p = 0.709). Thus, the low trial-by-trial evoked correla-
RSC values due to non-linear spike thresholds (de la Rocha et al., tions suggest that individual units respond independently from
2007). To assess this factor, we separately analyzed 2,658 unit each other on every trial. Indeed, within 50 ms after a foot shock
pairs with a geometric mean rate of <1 Hz and compared them a robust population response is easily observed, yet on average
to 506 unit pairs with a geometric mean rate that was similar to only 16% of units responded on each trial of a single foot shock
the awake spike rate (>1 Hz) in rats and monkeys (Aston-Jones (Figure 3). The proportion of neurons responding to the stimulus
and Bloom, 1981a; Bari and Aston-Jones, 2013; Bouret and in each trial remained relatively low when the post-stimulus win-
Sara, 2004; Foote et al., 1980; Kalwani et al., 2014; Takeuchi dow was increased to 100 ms (20% of units) or 200 ms (28% of
et al., 2016; Vankov et al., 1995). Only 19.2% of higher rate pairs units). Thus, the LC response to arousing stimuli in states char-
had significantly positive RSC values, which was similar to the acterized by slow waves (anesthesia and perhaps sleep) is not
percentage (15.8%) of lower rate pairs. Thus, correlated spiking necessarily a unitary population burst.
is not predominant between pairs of LC neurons, even when We further examined RSC between pairs of narrow or wide
firing rate is quantitatively similar to those observed in the awake units, as well as between pairs consisting of one narrow and
state. one wide unit (mixed). Correlations among each type of pair

Neuron 99, 1–14, September 5, 2018 3


Please cite this article in press as: Totah et al., The Locus Coeruleus Is a Complex and Differentiated Neuromodulatory System, Neuron (2018), https://
doi.org/10.1016/j.neuron.2018.07.037

Figure 2. RSC Values Were Near Zero and


Anti-correlations Related to Unit Waveform
Width
The RSC distribution was around zero (dotted blue
line) during spontaneous spiking (left panel) or
following single-pulse foot shock stimulation (right
panel). Distributions are shown for spontaneous
activity binned in 200 ms and for evoked activity in
a 50 ms window after the stimulus. Additional
windows were tested and yielded similar results
(see text). RSC values are plotted separately for
pairs of narrow units, pairs of wide units, and pairs
of mixed (narrow and wide) units according to the
labels on the left of the histograms. n = 100 (both
narrow units), 2,480 (both wide units), and 584
(mixed units) pairs. Prominent negative correla-
tions are apparent after evoked foot shocks
(arrows, right panel). These negative correlations
predominate in pairs containing a wide unit. See
also Figures S2 and S3.

and wide units were responsive to salient


stimuli, our results suggest a potential
local circuit architecture in which only
wide units generate sufficient somatic
NE release to cause local lateral
inhibition.

Synchrony Due to Putative Gap


Junctions or Common Synaptic
Input Is Rare
The presence of a minority of pairs with
highly positive RSC values suggests that
at least some LC single units are corre-
lated. We assessed the duration of coin-
cidental spiking between unit pairs by
measuring cross-correlograms between
spike trains. All cross-correlogram ana-
lyses used spontaneous spiking. We
were similar for spontaneous spiking (Welch’s F(2,3161) = 1.42, chose to initially study two timescales, tens of milliseconds
p = 0.245, u2 = 0.0003). Evoked correlations may differ by unit (‘‘broad-type’’ interactions) or sub-millisecond (‘‘sharp-type’’ in-
pair type, but the Kruskal-Wallis test was under-powered teractions), that could reflect either common synaptic input or
(H = 7.64, p = 0.022, u2 = 0.0003, power = 0.175). Post hoc tests gap junctions, respectively. Shared synaptic input from a third
suggest that pairs of mixed units may have a more negative neuron (or group of neurons) appears as a zero-centered peak
median correlation than pairs of wide-type units (p = 0.017). that is spread over tens of milliseconds (Perkel et al., 1967).
Nevertheless, the mean RSC values were near zero for all pair Gap junctions are associated with a sharp, sub-millisecond
types, which suggests that neither narrow nor wide units formed peak that is shifted 0.5 to 1 ms from zero (Trenholm et al.,
a highly correlated sub-population. 2014; van Welie et al., 2016). We observed coincident spiking
Strikingly, a large number of negative RSC values emerged dur- on both timescales (Figure 4A).
ing evoked activity. Furthermore, negatively correlated pairs Only 15% of pairs had significant cross-correlations (Fig-
were observed primarily when the pair included a wide unit (Fig- ure 4B). Of those correlated pairs, 62% had broad-type interac-
ure 2, arrows). Negative correlations, specifically after sensory tions, while 13% had sharp-type interactions, and the remaining
stimuli, may reflect lateral inhibition. Local noradrenergic inhibi- 25% had both broad- and sharp-type interactions. We also
tion is generated by somatic release of NE that occurs during observed a small number (0.6%) of interactions that were inhib-
the high frequency of spiking associated with sensory-evoked itory, but focus only on the excitatory ones.
responses (Huang et al., 2007), which inhibits neighboring neu- Pairs with broad-type interactions had higher RSC values in
rons via alpha-2 adrenoreceptors (Aghajanian et al., 1977; Ennis comparison to pairs with sharp-type interactions and pairs
and Aston-Jones, 1986; Lee et al., 1998). Given that both narrow without significant cross-correlations (Figure S4). These results

4 Neuron 99, 1–14, September 5, 2018


Please cite this article in press as: Totah et al., The Locus Coeruleus Is a Complex and Differentiated Neuromodulatory System, Neuron (2018), https://
doi.org/10.1016/j.neuron.2018.07.037

Figure 3. The Response to Single-Pulse


Foot Shock Engaged Different LC Units on
Each Trial
The figure shows an example recording of
20 units.
(A) The spike raster (trials are rows) after all
simultaneously recorded single-unit spike trains
were merged into a single multi-unit spike train per
rat. The red ticks indicate spiking during the first
50 ms after stimulation (i.e., the population phasic
burst of spikes). The phasic response is followed
by noradrenergic-mediated auto-inhibition. The
plot depicts 20 units from one rat.
(B) The trial-averaged peri-stimulus spike rate
histograms for single units were averaged. The
plot is the mean and SE across single units (n = 20
single units, same as in A, and bin size of 5 ms).
This plot shows that single units, on average
across trials, exhibit a phasic burst followed by
inhibition.
(C) The spike raster of the same 20 units (y axis) is
plotted for 500 ms before and 500 ms after foot
shock (arrows) on 10 randomly selected trials. The
spikes of each single unit that occur during the
50 ms window that encompasses the LC phasic
burst (in A and B) are colored in red. At the level of
single trials, different single units responded on
different trials. For example, unit 15 responded
during the 50 ms after foot shock only on trials 3, 5,
and 26, whereas unit 14 responded on trials 13
and 19. This randomness reduces the trial-by-trial
evoked RSC values.

are consistent with broad-type interactions and spike count LC neurons for 100 ms (Aghajanian et al., 1977), which corre-
correlations both reflecting common synaptic input. Broad- sponds with the delays observed here for broad-type interac-
type interactions, just like spike count correlations, did not tions (Figure 4D). Moreover, a comparable delay has been
depend on the distance between the unit pairs and therefore observed in paired intracellular recordings, which Alvarez
occurred with similar frequency throughout the LC nucleus (Fig- et al. (2002) demonstrated to be dependent upon the local LC
ure 4C). The distance invariance of correlated activity suggests activity that provides lateral inhibition (Alvarez et al., 2002; Ran-
an integration of broad, non-topographically organized afferent cic et al., 2018).
inputs. However, our findings of little correlated activity suggest Synchrony could conceivably also occur over multiple
the potentially synchronizing influence of shared synaptic input seconds or even minutes to hours, given that LC spiking
is somehow limited. is related to arousal (Aston-Jones and Bloom, 1981a). We
We next assessed the dynamics of broad-type interactions used a data-driven approach to detect synchrony occurring
by examining the peak times of the cross-correlograms for pairs in long windows ranging from 20 ms to 40 s (Bair et al.,
with significant interactions. In the example cross-correlograms 2001). Based on this analysis, we examined cross-correlo-
(Figure 4A), there is a notable diversity in the timing of the inter- grams over a ±20 s window. However, out of the already
action in different pairs. The interaction in Figure 4A1 was limited set of synchronous pairs observed, the vast majority
centered at 0 ms, while interactions between other pairs of synchronous spiking occurred in a window of 70 ms (Fig-
occurred before 0 ms (Figure 4A2) or after 0 ms (Figure 4A3). ure S5). Thus, the time windows we have explored throughout
Across the population of all pairs with significant broad-type in- these analyses are sensitive to the timescale of synchrony in
teractions, the cross-correlogram peak times were spread the LC.
over ±70 ms (Figure 4D). Therefore, neuronal interactions in
LC at this timescale may reflect common synaptic input in Sharp-Type Interactions Are Spatially Localized
cooperation with other mechanisms that introduce a delay. Only 6% of all recorded pairs had sharp-type interactions.
For example, lateral inhibition between units sharing a synaptic These sub-millisecond interactions fell off rapidly with the dis-
input could desynchronize their synaptic responses. The local tance between units, suggesting a dependence on electrotonic
NE release due to discharge of a single LC neuron (or a small coupling (Figure 4C). One possibility is that gap junctions
number of neurons) inhibits spontaneous spiking of neighboring spread synchrony throughout the LC using collections of

Neuron 99, 1–14, September 5, 2018 5


Please cite this article in press as: Totah et al., The Locus Coeruleus Is a Complex and Differentiated Neuromodulatory System, Neuron (2018), https://
doi.org/10.1016/j.neuron.2018.07.037

Figure 4. Spike Train Cross-correlograms


Indicated that Interactions Were Rare
(A) Three example cross-correlograms with sig-
nificant coincidental spiking on the timescale of
broad-type interactions (A1–A3) and sharp-type
interactions (A4–A6). An interaction was counted if
a pairwise 1% threshold (dotted gray lines) and a
1% global threshold (dotted blue lines) were
crossed. The mean of surrogate cross-correlo-
grams (solid gray line) was computed from jittered
spike times.
(B) A minority of cross-correlograms (15%) had a
significant number of coincidental spikes in at
least one bin. A larger percent of the pairs of nar-
row units had significant broad-type interactions in
comparison to pairs of wide units or mixed pairs.
(C) The percent of pairs with broad-type in-
teractions was similar regardless of the distance
between the units in the pair. On the other hand,
sharp-type interactions occurred only between
spatially confined units.
(D) Out of the pairs with significant broad-type
interactions, the timing of the peak of the cross-
correlogram is plotted as a black point and the
duration (continuous bins above the significance
threshold) is plotted as a gray line for each pair.
Timing and duration are presented as absolute
values (positive lags) because the sign of
the interaction is arbitrary. These interactions
occurred over a broad time range. Peaks were not
exclusively centered at time 0 (as in A1).
See also Figures S4 and S5.

electrotonically coupled neurons (Alvarez et al., 2002; Aston- Spiking of Individual LC Units Oscillates Asynchronously
Jones and Cohen, 2005; Usher et al., 1999). We counted the at Low (<2 Hz) Frequencies
number of units that exhibited sharp-type interactions with Synchrony could also emerge from entrainment with cortical
one or more other units. Out of the units with at least one slow rhythms, which are prominent during slow-wave sleep
sharp-type interaction with another unit, 38% interacted with and anesthesia (but also during the awake state) (Chan et al.,
only this one other unit (i.e., a network of 2 units), 35% inter- 2015; Clement et al., 2008; Lakatos et al., 2005; Leopold et al.,
acted with 2 other units, 16% interacted with 3 other units, 2003; Petersen et al., 2003; Sakata and Harris, 2012; Steriade
and the remaining 11% interacted with 4 to 6 other units. These et al., 1993). LC MUA has been shown to oscillate in slow
findings suggest that synchrony on the timescale of putative (1 Hz) frequency range and phase lock to the cortical slow
gap junctions is primarily limited to networks of 2 to 3 units, (1 Hz) oscillations, leading to the impression that the majority
but also as many as 7 units. Importantly, these data demon- of LC neurons spike in unison, entrained with the cortical rhythm
strate that (potentially gap junction-mediated) interacting LC (Eschenko et al., 2012; Lestienne et al., 1997; Safaai et al., 2015).
cells can form localized multi-neuron ensembles. We first revealed oscillations in the firing rate of LC single units
We assessed the propensity of these networks to produce by calculating the power spectrum of each unit’s spike train
repeating patterns of spiking over a few milliseconds (i.e., converted into a continuous spike density function (SDF). We
some unit A spikes would be consistently followed by unit B calculated the power spectrum of each single-unit SDF and
spikes around 1 ms later, which are followed by unit C spikes then examined the average power spectrum across all 234 single
around 1 ms later). We found that repeating patterns occurred units (Figure 5A). We observed peaks in the infra-slow fre-
with negligible frequency. In 2 out of 12 rats, we observed triplets quencies (Figure 5A), specifically at 0.09 Hz (periods of 11 s)
of units that spiked in a consistent order over 4 ms (allowing for and 0.4–0.5 Hz (periods of around 2 s), which were readily visible
0.4 ms jitter of each spike). Only 1 triplet out of 22,100 possible in SDFs (Figure 5G, top panel).
triplet patterns (0.005%) was found in one rat and 4 triplets Surprisingly, in contrast to the infra-slow (<1 Hz) range, we did
out of 1,330 possible triplet patterns (0.301%) were found not observe any distinct peak in the 1–2 Hz range. This result is
in the other rat. Patterns beyond triplets were never observed. unexpected in light of previous studies, which have found that LC
Sharp-type interactions (possibly reflecting electrotonic MUA oscillated around 1 Hz during anesthesia or slow-wave
coupling) are, therefore, spatially limited. sleep (Eschenko et al., 2012; Lestienne et al., 1997; Safaai

6 Neuron 99, 1–14, September 5, 2018


Please cite this article in press as: Totah et al., The Locus Coeruleus Is a Complex and Differentiated Neuromodulatory System, Neuron (2018), https://
doi.org/10.1016/j.neuron.2018.07.037

Figure 5. Spike Rates Oscillated Asynchro-


nously across Individual LC Units and
Narrow Units Preferred a Distinct Phase of
Cortical Slow Oscillations
(A) The plot shows the power spectrum averaged
across all single units. Two oscillatory regimes at
0.09 Hz and between 0.4 and 0.5 Hz were
observed in single-unit spike trains.
(B) Merging simultaneously recorded single-unit
spike trains into one multi-unit spike train revealed
spectral power around 1–2 Hz.
(C) Multi-unit spiking was phase locked to the
trough-to-peak transition reflected by the phase of
cortical delta oscillations. The polar plot is a his-
togram of the number of multi-units at each bin of
cortical LFP phase. The red line is the mean across
multi-units.
(D) Single units were also phase locked to the
cortical delta oscillation, despite not rhythmically
spiking at that frequency (see A). Narrow units
responded significantly earlier than wide units
during the cortical delta oscillation. The wide
units preferentially fired at 320 degrees, whereas
the mean angle for narrow units was 254 degrees
(the trough, or down state, was 180 degrees
and the zero-crossing between the trough and the
peak was at 270 degrees).
(E) Wide and narrow units both oscillated at infra-
slow frequencies (0.09 and 0.4–0.5 Hz), but narrow
unit spike trains had additional peaks of spectral
power between 0.1 and 0.2 Hz.
(F) The spike counts of narrow and wide units
fluctuated coherently at a range of infra-slow and
slow frequencies. The percent of pairs with sig-
nificant spike-spike coherence is plotted for pairs
of narrow units and pairs of wide units. Pairs of
wide units oscillated together at 0.09 and
0.4–0.5 Hz. Pairs of narrow units oscillated
together at different infra-slow frequencies and at
approximately 1.5 Hz.
(G) The spike rates of 3 example pairs over a 10 s
period are plotted as SDFs (top panel) and 0.4–
0.5 Hz bandpass filtered SDFs (middle panel). The
bottom panel is a histogram of the phase differ-
ences (4) between the units’ oscillations over the
entire recording session. The units in each pair
oscillated coherently throughout the recording,
but only the units in Pair 2 oscillated nearly syn-
chronously (i.e., with a mean 4 of 0 degrees). The
mean 4 for each example pair is marked by the red
line on the polar plot.
(H) At the population level, the mean 4 values are
distributed uniformly.
See also Figure S6.

et al., 2015). To understand the relationship between the periodic delta oscillations (Figure 5C). Our results demonstrate that if only
firing of LC single units and slow fluctuations of cortical activity, multi-unit spiking is measured (as is typical in LC recordings), the
we first compared our results with prior studies of LC MUA by data suggest that LC neurons respond synchronously along with
merging all simultaneously recorded spike trains into a single cortical slow rhythms; however, our data reveal that this is not
multi-unit spike train and converting that to an SDF. In line with the case at the single-unit level. In spite of LC single units not
previous studies, we observed that LC MUA oscillated in the exhibiting spike rate fluctuations at 1–2 Hz (Figure 5A), approx-
delta frequency band (Figure 5B). Out of eight rats with spiking imately 69% of single units were significantly phase locked to the
during cortical delta oscillations, all LC MUA signals were signif- cortical delta (Rayleigh’s test for circular uniformity, p < 0.05).
icantly phase locked (Rayleigh’s Z test, p < 0.05) to cortical LFP Individual LC neurons, therefore, respond on different delta

Neuron 99, 1–14, September 5, 2018 7


Please cite this article in press as: Totah et al., The Locus Coeruleus Is a Complex and Differentiated Neuromodulatory System, Neuron (2018), https://
doi.org/10.1016/j.neuron.2018.07.037

cycles rather than as a synchronized, rhythmically fluctuating sory stimuli, in line with our earlier pairwise analysis (Figures
population (which would yield a peak at 1–2 Hz in the single- 2B, 2C, and 3).
unit power spectrum). We next attempted to detect and discriminate which LC units
Similar proportions of each unit type (70% of wide units and formed correlated sub-populations spiking together as ensem-
66% of narrow units) were phase locked to cortical LFP delta bles using graph theory analysis, which were detected based
(two-sided Fisher’s exact test, odds ratio = 1.23, CI = [0.470 on RSC values (STAR Methods). We observed ensembles in
3.202], p = 0.803). Notably, narrow units responded significantly each set of simultaneously recorded units (Figure 6E). We iden-
earlier in the cortical delta cycle (Watson-Williams test for equal tified a total of 23 ensembles, ranging from 1 to 3 per rat, and
circular means, F(83) = 40.959, p < 0.0001; Figure 5D). In consisting of 2 to 9 units per ensemble. Ensembles were most
contrast to the canonical thinking that LC neurons respond likely due to distance-invariant shared synaptic inputs (Figures
homogenously during a particular phase of cortical slow waves, S2 and 4C), which contributed the majority of correlated activity
our work demonstrates that subsets of LC cells differing by in the nucleus; correspondingly, LC unit ensembles were
waveform width (and other physiological properties) also have spatially diffuse (Figure S7). This contrasts the highly localized
differential preferred phases of cortical slow activity. networks observed for sub-millisecond (putative gap junction)
Peaks in the spike rate power spectra were apparent in the sharp-type interactions. Strikingly, the units in an ensemble pre-
infra-slow range at 0.09 and 0.4–0.5 Hz (Figure 5A). Both narrow dominantly consisted of either wide or narrow units (Figure 6F).
and wide unit spike rates oscillated at these frequencies and nar-
row units also oscillated at additional frequencies between 0.1 A Minority of Correlated Single Units Provide Targeted
and 0.2 Hz (Figure 5E). These peaks reflect rhythmic fluctuations Forebrain Neuromodulation
in spike rate that are predominant in many single units, but not We examined the degree to which correlated units have overlap-
necessarily synchronous across units. To assess the synchrony ping projection targets. We assessed the projection properties of
of infra-slow spike rate oscillations, we assessed spike-spike LC cells by applying direct electrical stimulation in up to 15 fore-
coherence and observed strong pairwise coherence (Figure 5F), brain sites. Example spike rasters showing antidromic spikes,
which suggests that cell pairs have a stable oscillatory relation- response latencies, number of projection targets, and firing rates
ship, although the timing of spike rate increases may not be of units based on projection target are shown in Figure S8. The
in-phase. Therefore, we next examined the phase relationship mean latencies for each projection target are consistent with
(4) of the spike rate oscillations between LC units in the infra- the prior literature (Jodo et al., 1998; Nakamura, 1977; Nakamura
slow range. Spiking of the majority of pairs (73% for 0.4–0.5 Hz and Iwama, 1975). In general, positively or negatively correlated
and 67% for 0.09 Hz) oscillated coherently with a stable phase unit pairs (those with significant correlation at p < 0.01, permuta-
relationship (Rayleigh’s test for circular uniformity, p < 0.05). tion test) did not have any greater tendency for both units to
Examples for coherent spike rate oscillations at 0.4 to 0.5 Hz jointly project to overlapping forebrain targets (Figures 7A and
are shown in Figure 5G and additional examples are shown in 7B). Additionally, pairs with broad-type interactions (assessed
Figure S6. At the population level (3,164 pairs; Figure 5H), by spike train cross-correlograms) did not relate to the degree
the mean 4 values were distributed uniformly for spike rate of target overlap between units (Figure 7C). However, pairs
oscillations at both 0.09 (Rayleigh’s Z = 2.531, p = 0.080) and with sharp-type interactions were more likely to project to the
0.4–0.5 Hz (Rayleigh’s Z = 1.074, p = 0.342). These data indicate same target (Figure 7D). More pairs with sharp-type interactions
that periodic (infra-slow) elevations in spike rate are synchro- (88%) than non-interacting cell pairs (72%) had overlapping pro-
nized among subsets of neurons, but those elevations are out jections to the same forebrain zone (Cohen’s D = 0.55, one-sided
of phase, or even anti-phase, with other subsets of neurons. At Fisher’s exact test, p = 0.075). Out of the unit pairs with overlap-
the population level, this finding demonstrated ensembles con- ping projection zones (e.g., any division of prefrontal cortex or
structed from seconds-long oscillatory rate changes. sensory cortex or thalamus), cell pairs had a similar tendency
to project to cortex more than thalamus and prefrontal cortex
LC Single Units Exhibit Complex Population Patterns over sensory cortex (Figure 7D, inset).
and Form Ensembles
Analysis of cell pairs has, thus far, suggested that LC ensembles DISCUSSION
may exist. We explored ensembles more directly by moving to
beyond-pairwise measures. We assessed synchrony between The LC Is Capable of an Ensemble-Based Code
single neurons and the surrounding population of single units The degree of population synchrony in the LC, its timescales,
(‘‘population coupling’’) using the method of Okun et al. (2015) spatial distribution, and relation to projection targets are crucial
(STAR Methods). During spontaneous activity, population to understand neuromodulation by this diffusely projecting
coupling varied (Figure 6B), suggesting that some subsets of nucleus. Emerging research on individually recorded LC single
units may be synchronously active as ensembles. units has hinted at the possibility of one-to-one and ensemble-
Sensory stimulation is thought to evoke synchronous based connection schemes (Chandler et al., 2014; Hirschberg
discharge of many LC neurons (Aston-Jones and Bloom, et al., 2017; Uematsu et al., 2017). Here, we examined this issue
1981b) and may therefore result in strong population coupling. systematically with the first large-scale simultaneous recordings
However, foot shocks did not cause a large number of single of many single units. We demonstrate an ensemble code in the
units to couple with the population (Figures 6C and 6D), suggest- LC. Our main finding was that, during anesthesia, LC population
ing a lack of totally synchronized population discharge to sen- dynamics were characterized by sparse, dynamically regulated

8 Neuron 99, 1–14, September 5, 2018


Please cite this article in press as: Totah et al., The Locus Coeruleus Is a Complex and Differentiated Neuromodulatory System, Neuron (2018), https://
doi.org/10.1016/j.neuron.2018.07.037

Figure 6. LC Single Units Have Diverse


Coupling to the Population and Organize
into Ensembles of Units with the Same
Waveform Characteristics
(A) Population coupling was calculated as the
cross-correlogram between each single unit and
the merged spike train of all remaining single units
that were simultaneously recorded. Example unit
A illustrates a case of strong population coupling.
Spikes of unit A coincided with spikes of many
other units recorded simultaneously. Example unit
B illustrates a lack of population coupling.
(B) A histogram of the population coupling
strength (Z score at time 0) illustrates the broad
range of population coupling strength across LC
single units. Dotted lines show a Z score of 2 for
reference.
(C and D) Population coupling using spikes during
the 200 ms after a single foot shock (C) or after a
brief train (5 mA pulses at 30 Hz) of foot shocks (D).
As with spontaneous activity, many single units
are not coupled to the population, with the
exception of those units on the right tail of the
distribution.
(E) Examples of ensembles detected in two rats.
White dots indicate correlated units. Simulta-
neously recorded single units were treated as a
network with links between correlated pairs and
ensembles were detected using community
detection algorithms. Magenta lines outline en-
sembles.
(F) The percent of each unit (narrow or wide)
making up each ensemble. Each bar is one of 23
ensembles. The majority of units in each ensemble
had the same waveform characteristics.
See also Figure S7.

correlations among subsets of multiple (more than two) single awake LC activity (>1 Hz) or studied evoked firing (which
units over briefer (sub-millisecond), intermediate (hundreds of elicits higher spike rates), the proportion of correlated pairs re-
milliseconds), and long (seconds) periodic (delta and infra- mained similarly small. In sum, minor suppression of firing rate
slow) timescales. Our work reveals that spiking under urethane by urethane is likely to account for only a small reduction in
anesthesia is not an en masse population event (RSC was synchrony. In addition to the aforementioned potentially sup-
0.04, only 15% of cross-correlograms indicated synchronous pressive effects of anesthesia on LC synchrony, it is also
spiking, and sensory stimuli evoked a synchronized response possible that anesthesia could increase LC synchrony. For
from only 16% of unit pairs). example, in the cortex, anesthesia is associated with both
Urethane anesthesia may have influenced synchrony within slow EEG/LFP oscillations and higher RSC between cortical
the LC in a number of ways. It may have suppressed LC syn- neurons (Aasebø et al., 2017; Ecker et al., 2014; Erchova
chrony due to its depressive effects on spike rate (mediated et al., 2002; Greenberg et al., 2008; Sakata and Harris,
by weak effects on synaptic and non-synaptic currents), which 2009). The higher cortical RSC results from the ‘‘temporal com-
is associated with a reduced RSC (Daló and Hackman, 2013; pacting’’ of spiking from many cortical neurons into windows
de la Rocha et al., 2007; Sceniak and Maciver, 2006). How- of increased population activity. Given that the cortex and
ever, the spike rate reductions in the LC are minor under anes- LC are reciprocally connected, an increase in cortical network
thesia (0.89 Hz versus means ranging from 0.92 to 1.4 Hz in synchrony during anesthesia would be expected to increase
the awake rat and monkey) (Aston-Jones and Bloom, 1981a; correlations among at least some LC afferents. This factor
Bari and Aston-Jones, 2013; Bouret and Sara, 2004; Foote could increase LC synchrony because increasing correlations,
et al., 1980; Kalwani et al., 2014; Takeuchi et al., 2016; Vankov even among some inputs to a network (i.e., inputs to LC in this
et al., 1995). Given that higher firing rates lead, in general, to case), will increase correlations in that network (Doiron et al.,
higher values of pairwise correlations due to the non-linearities 2016). In support of this view, we demonstrate that spontane-
of spike thresholds (de la Rocha et al., 2007), such minor firing ously occurring states with higher cortical synchrony (high-
rate reductions are likely to account for little reduction in amplitude slow LFP oscillations) are associated with higher
correlated activity. Moreover, when we considered correla- RSC values in LC. It is important, however, to note that the
tions among only pairs with firing rates with values typical of LC receives inputs from many structures and it is not possible

Neuron 99, 1–14, September 5, 2018 9


Please cite this article in press as: Totah et al., The Locus Coeruleus Is a Complex and Differentiated Neuromodulatory System, Neuron (2018), https://
doi.org/10.1016/j.neuron.2018.07.037

Figure 7. The Minority of Unit Pairs with


Synchrony on the Timescale of Gap Junc-
tions Provided Targeted Forebrain Neuro-
modulation
(A) RSC values were divided into highly correlated
pairs (Pearson’s correlation coefficient, p < 0.001)
with positive (blue) or negative (purple) correla-
tions.
(B) The percent of pairs with both units jointly
projecting to the same forebrain target did not
differ between pairs with correlated activity (blue),
anti-correlated activity (purple), and non-corre-
lated units (gray). The percent is out of the total
number of units indicated in the figure legend.
Targets were defined as either individual brain
regions or as zones (i.e., cortical, sub-cortical,
thalamic, prefrontal, primary sensory cortex, or
secondary sensory cortex). Zones were examined
because prior work has indicated that single LC
neurons may project to multiple functionally
related forebrain sites (Simpson et al., 1997).
(C) The percent of pairs with overlapping projec-
tion targets did not depend on the pair having a
network interaction.
(D) Pairs with gap junction interactions had a
significantly greater likelihood of projecting to the
same forebrain target zone.
See also Figure S8.

to directly infer that an increase in cortical synchrony during Putative Mechanisms Underlying LC Ensemble Activity
anesthesia implies that LC synchrony is higher during anes- Prior accounts have suggested that gap junctions could
thesia, as well. In sum, it is unknown if anesthesia may have contribute to synchrony in LC (Alvarez et al., 2002; Aston-Jones
increased synchrony in the LC, as it does in cortex; however, and Cohen, 2005; Usher et al., 1999) given their presence in the
anesthesia seems unlikely to have suppressed LC synchrony developing and adult rat LC (Alvarez et al., 2002; Rash et al.,
through reduced firing rates or, if so, then not by a large 2007; Van Bockstaele et al., 2004). Sharp, sub-millisecond inter-
amount. actions are likely due to gap junctions (Trenholm et al., 2014; van
We demonstrated that synchrony was consistently sparse Welie et al., 2016). We found such interactions coupling small
within the LC under anesthesia. However, the results of an networks of LC neurons, although synchrony did not unfold in
earlier study in the awake monkey (Usher et al., 1999) suggest a succession of spikes beyond two units, suggesting that such
that a variety of degrees of LC synchrony possibly occur across brief synchrony is limited and spatially confined. We also
other brain and behavioral states. Despite a substantial differ- observed cross-correlogram peaks over tens of milliseconds,
ence in the proportion of correlated cell pairs (80% versus which were delayed by 10 to 70 ms. We propose that this
15% in our study), the timescale (100 ms) of interactions is delayed synchrony is due to converging effects of shared synap-
consistent across both studies. Moreover, we assessed spike tic inputs (extra-LC) and lateral inhibition (intra-LC). LC cells with
train cross-correlograms over a large range (0.5 ms up to 40 s) wide waveforms may be a crucial controller of LC synchrony,
and found that LC correlated activity was predominant in the given our findings that these units may be the source of lateral
timescale of 100 ms, which also agrees with in vitro results inhibition.
(Alvarez et al., 2002; Rancic et al., 2018). Additionally, we
observed brief (sub-millisecond) interactions between cell pairs Potential for Targeted Neuromodulation by LC
and ensembles. Our results further extend the current knowl- Ensembles
edge about LC cell interactions in the spatial domain. We Rather than a purely non-specific neuromodulatory system, our
show the distance dependence of pairwise and ensemble data suggest a few possibilities for targeted noradrenergic
synchrony and the lack of successive sharp-type interactions forebrain neuromodulation. We showed that putatively gap junc-
beyond two nearby cells. Thus, previous findings that LC syn- tion coupled pairs were more likely to project to similar forebrain
chrony may be high during vigilant states (Usher et al., 1999) regions. Targeted forebrain neuromodulation may also be
and the evidence for an ensemble-based code, reported here, achieved by NE gradients between forebrain sites, which could
suggest that ensemble-based signaling from LC could modulate be generated by infra-slow (<1 Hz) oscillations in spike rate.
various forebrain networks associated with different brain and Our results demonstrate that LC neurons will have in-phase
behavioral states. spike rate oscillations with other neurons (forming an ensemble)

10 Neuron 99, 1–14, September 5, 2018


Please cite this article in press as: Totah et al., The Locus Coeruleus Is a Complex and Differentiated Neuromodulatory System, Neuron (2018), https://
doi.org/10.1016/j.neuron.2018.07.037

and out-of-phase or anti-phase spike rate oscillations with other may correspond to those LC units involved in analgesia. Our
neurons (thus excluding those neurons from the ensemble). work reveals a new level of diversity among LC single units
Accordingly, over an infra-slow duration (2 to 10 s), some LC that introduces many outstanding questions, from the local cir-
neurons will synchronously release NE to their projection targets, cuit and afferent connectivity principles that underlie formation
while other LC neurons are suppressed. Depending on their of LC ensembles among cells with similar waveform widths, to
projection targets (which can often be single forebrain regions how those ensemble activity patterns are related to cortical LFP.
according to our data and others; Kebschull et al., 2016), NE
could be simultaneously high in some forebrain regions and Implications for Understanding the Noradrenergic
simultaneously low in others over a timescale of 2 to 10 s. Neuromodulatory System
It is important to bear in mind that a great deal of evidence
LC Cells Differing in Waveform Width Present supports the notion of global noradrenergic neuromodulation
Functionally Differentiated Features and our findings do not contradict this view, but rather refine it.
We identified two types of LC units that were clustered by wave- Volume NE release provides simultaneous neurotransmission
form width. Remarkably, waveform width was associated with a in distant brain regions on a timescale of a few seconds (As-
number of distinct functional properties: (1) divergent spatial ton-Jones et al., 1980; Mundorf et al., 2001). This global and rela-
clustering (Figure 1E), (2) different firing rates (Figure 1C), (3) con- tively slow neuromodulation is clearly synchronous by nature and
trasting propensities for possible NE-mediated lateral inhibition presents a critical component of controlling brain state, espe-
(Figure 2), (4) distinct phase preferences for cortical slow LFP cially on timescales relevant to arousal and behavior. By demon-
oscillation (Figure 5D), (5) spike rate oscillations at different fre- strating LC ensembles active over a range of timescales, our
quencies (Figure 5E), (6) spike rate oscillations were coherent at results support the possibility that particular combinations of
different frequencies of spike-spike coherence for pairs of unit activations may provide more targeted forebrain modulation.
narrow units versus wide units (Figure 5F), and (7) predominant Moreover, LC cells have diverse physiological and functional
formation of cell ensembles from LC units with the same wave- properties. These features may allow a more nuanced role for
form width (Figure 6F). It remains to be seen if there is an under- the LC in theories of systems and cognitive neuroscience.
lying genetic, morphological, or membrane physiological relation-
ship between waveform widths and the functional properties of STAR+METHODS
neurons. In particular, further investigation of the relationship be-
tween cortical LFP and LC cells with differing waveforms during Detailed methods are provided in the online version of this paper
natural states of sleep and wakefulness is warranted. The delta and include the following:
frequency band has similar features in the awake, sleeping, and
anesthetized rodent (Clement et al., 2008; Bermudez Contreras d KEY RESOURCES TABLE
et al., 2013; Petersen et al., 2003; Sakata and Harris, 2012). The d CONTACT FOR REAGENT AND RESOURCE SHARING
phase preferences of narrow and wide LC units may be relevant d EXPERIMENTAL MODEL AND SUBJECT DETAILS
to slow-wave activity in sleep and wakefulness. B Subjects
The physiological properties that give rise to narrow and wide d METHOD DETAILS
waveform widths await additional study; however, it is unlikely B Anesthesia and Surgical Procedures
that the differences in waveform width are due to electrode posi- B Stereotaxic coordinates and electrode placement
tioning. The primary determinant of the width of the extracellu- B Electrodes
larly recorded action potential is the soma, not the dendritic B Recording and signal acquisition
morphology of neurons (Gold et al., 2006). The soma in the LC B Characterization of LC units
core are highly similar, except for a minority of small round cells B Sensory stimuli
that are found ventrally (Loughlin et al., 1986). However, the B Intra-cranial stimulation
ventrally distributed narrow units are likely not these small round B Administration of clonidine
cells because such cells project only to hypothalamus or do not B Histology
project out of the nucleus (Cintra et al., 1982; Loughlin et al., d QUANTIFICATION AND STATISTICAL ANALYSIS
1986), whereas the recorded narrow units projected to B Spike detection
numerous forebrain regions with similar frequency as wide units. B Spike clustering
The predominant feature of narrow units was a brief trough. B Detection of spikes across tetrodes
Increasing electrode distance can widen the spike trough by B Assigning unit location on recording array
up to 0.2 ms (Gold et al., 2006), suggesting that narrow units B Spike count correlations
could be an artifact of higher proximity to the electrode; however, B Cross-correlograms
it is unlikely that electrode distance from the soma accounted for B Syn-fire chain analysis
the large trough width difference between narrow and wide units B Population coupling
of 0.6 ms. Moreover, if narrow units were closer to the electrode, B Graph theory analysis and ensemble detection
then they would have larger amplitudes than wide units, but this B Oscillations in spike count
was not the case. The ventral LC core projects to the spinal cord B Spike-LFP phase locking
and hindbrain (Li et al., 2016; Loughlin et al., 1982; Schwarz et al., B Antidromic spiking
2015) and narrow units, with their distinct functional properties, d DATA AND SOFTWARE AVAILABILITY

Neuron 99, 1–14, September 5, 2018 11


Please cite this article in press as: Totah et al., The Locus Coeruleus Is a Complex and Differentiated Neuromodulatory System, Neuron (2018), https://
doi.org/10.1016/j.neuron.2018.07.037

SUPPLEMENTAL INFORMATION Bair, W., Zohary, E., and Newsome, W.T. (2001). Correlated firing in macaque
visual area MT: time scales and relationship to behavior. J. Neurosci. 21,
Supplemental Information includes ten figures and one table and can be found 1676–1697.
with this article online at https://doi.org/10.1016/j.neuron.2018.07.037. Bari, A., and Aston-Jones, G. (2013). Atomoxetine modulates spontaneous
and sensory-evoked discharge of locus coeruleus noradrenergic neurons.
ACKNOWLEDGMENTS Neuropharmacology 64, 53–64.
Bellesi, M., Tononi, G., Cirelli, C., and Serra, P.A. (2016). Region-specific
We thank Axel Oeltermann, Eduard Krampe, and Joachim Werner for technical dissociation between cortical noradrenaline levels and the sleep/wake cycle.
assistance; Dr. Henry Evrard for expertise in immunohistochemistry and Sleep 39, 143–154.
Marcel Hertl, Felicitas Horn, Jennifer Smuda, and Katalin Kalya for assistance
Berens, P. (2009). CircStat: a MATLAB toolbox for circular statistics. J. Stat.
with histology; and Valeria d’Andrea for helping to write code for the permuta-
Softw. 31, 1–21.
tion test. We thank Dr. Susan Sara and Dr. Tony Pickering for comments on the
paper. This research was supported by European Union FP7 funding as a Bermudez Contreras, E.J., Schjetnan, A.G., Muhammad, A., Bartho, P.,
Marie Curie Fellowship to N.K.T. (PIIF-GA-2012-331122) and FET Open fund- McNaughton, B.L., Kolb, B., Gruber, A.J., and Luczak, A. (2013). Formation
ing to S.P., O.E., and N.K.L. (SICODE). and reverberation of sequential neural activity patterns evoked by sensory
stimulation are enhanced during cortical desynchronization. Neuron 79,
555–566.
AUTHOR CONTRIBUTIONS
Bouret, S., and Sara, S.J. (2004). Reward expectation, orientation of attention
Conceptualization, N.K.T., N.K.L., and O.E.; Methodology, N.K.T., O.E., and and locus coeruleus-medial frontal cortex interplay during learning. Eur. J.
S.P.; Formal Analysis, N.K.T.; Investigation, N.K.T. and R.M.N.; Resources, Neurosci. 20, 791–802.
N.K.L.; Writing – Original Draft, N.K.T.; Writing – Review & Editing, N.K.T.,
Bullmore, E., and Sporns, O. (2009). Complex brain networks: graph theoret-
S.P., N.K.L., and O.E.; Visualization, N.K.T.; Supervision, O.E.; Funding Acqui-
ical analysis of structural and functional systems. Nat. Rev. Neurosci. 10,
sition, N.K.L. and S.P.
186–198.
Cedarbaum, J.M., and Aghajanian, G.K. (1978). Activation of locus coeruleus
DECLARATION OF INTERESTS
neurons by peripheral stimuli: modulation by a collateral inhibitory mechanism.
The authors declare no competing interests. Life Sci. 23, 1383–1392.
Chan, A.W., Mohajerani, M.H., LeDue, J.M., Wang, Y.T., and Murphy, T.H.
Received: December 7, 2017 (2015). Mesoscale infraslow spontaneous membrane potential fluctuations
Revised: March 25, 2018 recapitulate high-frequency activity cortical motifs. Nat. Commun. 6, 7738.
Accepted: July 20, 2018 Chandler, D.J., Gao, W.-J., and Waterhouse, B.D. (2014). Heterogeneous
Published: August 16, 2018 organization of the locus coeruleus projections to prefrontal and motor
cortices. Proc. Natl. Acad. Sci. USA 111, 6816–6821.
SUPPORTING CITATIONS
Chen, F.-J., and Sara, S.J. (2007). Locus coeruleus activation by foot shock or
electrical stimulation inhibits amygdala neurons. Neuroscience 144, 472–481.
The following references appear in the Supplemental Information: Bellesi et al.
(2016); Morrison et al. (1979). Christie, M.J., Williams, J.T., and North, R.A. (1989). Electrical coupling
synchronizes subthreshold activity in locus coeruleus neurons in vitro from
neonatal rats. J. Neurosci. 9, 3584–3589.
REFERENCES
Cintra, L., Dı́az-Cintra, S., Kemper, T., and Morgane, P.J. (1982). Nucleus locus
Aasebø, I.E.J., Lepperød, M.E., Stavrinou, M., Nøkkevangen, S., Einevoll, G., coeruleus: a morphometric Golgi study in rats of three age groups. Brain Res.
Hafting, T., and Fyhn, M. (2017). Temporal processing in the visual cortex of the 247, 17–28.
awake and anesthetized rat. eNeuro 4, https://doi.org/10.1523/ENEURO. Clement, E.A., Richard, A., Thwaites, M., Ailon, J., Peters, S., and Dickson,
0059-17.2017. C.T. (2008). Cyclic and sleep-like spontaneous alternations of brain state
Aghajanian, G.K., Cedarbaum, J.M., and Wang, R.Y. (1977). Evidence for under urethane anaesthesia. PLoS ONE 3, e2004.
norepinephrine-mediated collateral inhibition of locus coeruleus neurons. Cohen, M.R., and Kohn, A. (2011). Measuring and interpreting neuronal corre-
Brain Res. 136, 570–577. lations. Nat. Neurosci. 14, 811–819.
Alvarez, V.A., Chow, C.C., Van Bockstaele, E.J., and Williams, J.T. (2002). Daló, N.L., and Hackman, J.C. (2013). The anesthetic urethane blocks excit-
Frequency-dependent synchrony in locus ceruleus: role of electrotonic atory amino acid responses but not GABA responses in isolated frog spinal
coupling. Proc. Natl. Acad. Sci. USA 99, 4032–4036. cords. J. Anesth. 27, 98–103.
Aston-Jones, G., and Bloom, F.E. (1981a). Activity of norepinephrine-contain- Dayan, P., and Yu, A.J. (2006). Phasic norepinephrine: a neural interrupt signal
ing locus coeruleus neurons in behaving rats anticipates fluctuations in the for unexpected events. Network 17, 335–350.
sleep-waking cycle. J. Neurosci. 1, 876–886. , K., and Reyes, A. (2007).
de la Rocha, J., Doiron, B., Shea-Brown, E., Josic
Aston-Jones, G., and Bloom, F.E. (1981b). Norepinephrine-containing locus Correlation between neural spike trains increases with firing rate. Nature
coeruleus neurons in behaving rats exhibit pronounced responses to non- 448, 802–806.
noxious environmental stimuli. J. Neurosci. 1, 887–900. Devilbiss, D.M., and Berridge, C.W. (2006). Low-dose methylphenidate
Aston-Jones, G., and Cohen, J.D. (2005). An integrative theory of locus coeru- actions on tonic and phasic locus coeruleus discharge. J. Pharmacol. Exp.
leus-norepinephrine function: adaptive gain and optimal performance. Annu. Ther. 319, 1327–1335.
Rev. Neurosci. 28, 403–450. , K.
Doiron, B., Litwin-Kumar, A., Rosenbaum, R., Ocker, G.K., and Josic
Aston-Jones, G., Segal, M., and Bloom, F.E. (1980). Brain aminergic (2016). The mechanics of state-dependent neural correlations. Nat.
axons exhibit marked variability in conduction velocity. Brain Res. 195, Neurosci. 19, 383–393.
215–222. Ecker, A.S., Berens, P., Cotton, R.J., Subramaniyan, M., Denfield, G.H.,
Aston-Jones, G., Foote, S.L., and Segal, M. (1985). Impulse conduction prop- Cadwell, C.R., Smirnakis, S.M., Bethge, M., and Tolias, A.S. (2014). State
erties of noradrenergic locus coeruleus axons projecting to monkey cerebro- dependence of noise correlations in macaque primary visual cortex. Neuron
cortex. Neuroscience 15, 765–777. 82, 235–248.

12 Neuron 99, 1–14, September 5, 2018


Please cite this article in press as: Totah et al., The Locus Coeruleus Is a Complex and Differentiated Neuromodulatory System, Neuron (2018), https://
doi.org/10.1016/j.neuron.2018.07.037

Engel, A.K., König, P., Gray, C.M., and Singer, W. (1990). Stimulus-dependent Leopold, D.A., Murayama, Y., and Logothetis, N.K. (2003). Very slow activity
neuronal oscillations in cat visual cortex: inter-columnar interaction as deter- fluctuations in monkey visual cortex: implications for functional brain imaging.
mined by cross-correlation analysis. Eur. J. Neurosci. 2, 588–606. Cereb. Cortex 13, 422–433.
Ennis, M., and Aston-Jones, G. (1986). Evidence for self- and neighbor-medi- Lestienne, R., Hervé-Minvielle, A., Robinson, D., Briois, L., and Sara, S.J.
ated postactivation inhibition of locus coeruleus neurons. Brain Res. 374, (1997). Slow oscillations as a probe of the dynamics of the locus coeruleus-
299–305. frontal cortex interaction in anesthetized rats. J. Physiol. Paris 91, 273–284.

Erchova, I.A., Lebedev, M.A., and Diamond, M.E. (2002). Somatosensory Li, Y., Hickey, L., Perrins, R., Werlen, E., Patel, A.A., Hirschberg, S., Jones,
cortical neuronal population activity across states of anaesthesia. Eur. J. M.W., Salinas, S., Kremer, E.J., and Pickering, A.E. (2016). Retrograde optoge-
Neurosci. 15, 744–752. netic characterization of the pontospinal module of the locus coeruleus with a
canine adenoviral vector. Brain Res. 1641 (Pt B), 274–290.
Eschenko, O., Magri, C., Panzeri, S., and Sara, S.J. (2012). Noradrenergic neu-
rons of the locus coeruleus are phase locked to cortical up-down states during Loughlin, S.E., Foote, S.L., and Fallon, J.H. (1982). Locus coeruleus projec-
sleep. Cereb. Cortex 22, 426–435. tions to cortex: topography, morphology and collateralization. Brain Res.
Bull. 9, 287–294.
Finlayson, P.G., and Marshall, K.C. (1988). Synchronous bursting of locus
Loughlin, S.E., Foote, S.L., and Grzanna, R. (1986). Efferent projections of
coeruleus neurons in tissue culture. Neuroscience 24, 217–225.
nucleus locus coeruleus: morphologic subpopulations have different efferent
Foote, S.L., Aston-Jones, G., and Bloom, F.E. (1980). Impulse activity of locus targets. Neuroscience 18, 307–319.
coeruleus neurons in awake rats and monkeys is a function of sensory stimu-
Luongo, F.J., Zimmerman, C.A., Horn, M.E., and Sohal, V.S. (2016).
lation and arousal. Proc. Natl. Acad. Sci. USA 77, 3033–3037.
Correlations between prefrontal neurons form a small-world network that op-
Fujisawa, S., Amarasingham, A., Harrison, M.T., and Buzsáki, G. (2008). timizes the generation of multineuron sequences of activity. J. Neurophysiol.
Behavior-dependent short-term assembly dynamics in the medial prefrontal 115, 2359–2375.
cortex. Nat. Neurosci. 11, 823–833.
Mana, M.J., and Grace, A.A. (1997). Chronic cold stress alters the basal and
Gold, C., Henze, D.A., Koch, C., and Buzsáki, G. (2006). On the origin of the evoked electrophysiological activity of rat locus coeruleus neurons.
extracellular action potential waveform: a modeling study. J. Neurophysiol. Neuroscience 81, 1055–1064.
95, 3113–3128. McCune, S.K., Voigt, M.M., and Hill, J.M. (1993). Expression of multiple alpha
Greenberg, D.S., Houweling, A.R., and Kerr, J.N.D. (2008). Population imaging adrenergic receptor subtype messenger RNAs in the adult rat brain.
of ongoing neuronal activity in the visual cortex of awake rats. Nat. Neurosci. Neuroscience 57, 143–151.
11, 749–751. Mitchell, K., Oke, A.F., and Adams, R.N. (1994). In vivo dynamics of norepi-
Grzanna, R., and Molliver, M.E. (1980). The locus coeruleus in the rat: an immu- nephrine release-reuptake in multiple terminal field regions of rat brain.
nohistochemical delineation. Neuroscience 5, 21–40. J. Neurochem. 63, 917–926.

Hirschberg, S., Li, Y., Randall, A., Kremer, E.J., and Pickering, A.E. (2017). Mitra, P., and Bokil, H. (2007). Observed Brain Dynamics (Oxford
Functional dichotomy in spinal- vs prefrontal-projecting locus coeruleus mod- University Press).
ules splits descending noradrenergic analgesia from ascending aversion and Morrison, J.H., Molliver, M.E., and Grzanna, R. (1979). Noradrenergic innerva-
anxiety in rats. eLife 6, 570. tion of cerebral cortex: widespread effects of local cortical lesions. Science
205, 313–316.
Huang, H.-P., Wang, S.-R., Yao, W., Zhang, C., Zhou, Y., Chen, X.-W., Zhang,
B., Xiong, W., Wang, L.-Y., Zheng, L.-H., et al. (2007). Long latency of evoked Mundorf, M.L., Joseph, J.D., Austin, C.M., Caron, M.G., and Wightman, R.M.
quantal transmitter release from somata of locus coeruleus neurons in rat (2001). Catecholamine release and uptake in the mouse prefrontal cortex.
pontine slices. Proc. Natl. Acad. Sci. USA 104, 1401–1406. J. Neurochem. 79, 130–142.

Ikegaya, Y., Aaron, G., Cossart, R., Aronov, D., Lampl, I., Ferster, D., and Nakamura, S. (1977). Some electrophysiological properties of neurones of rat
Yuste, R. (2004). Synfire chains and cortical songs: temporal modules of locus coeruleus. J. Physiol. 267, 641–658.
cortical activity. Science 304, 559–564. Nakamura, S., and Iwama, K. (1975). Antidromic activation of the rat locus co-
eruleus neurons from hippocampus, cerebral and cerebellar cortices. Brain
Ishimatsu, M., and Williams, J.T. (1996). Synchronous activity in locus coeru-
Res. 99, 372–376.
leus results from dendritic interactions in pericoerulear regions. J. Neurosci.
16, 5196–5204. Neves, R.M., van Keulen, S., Yang, M., Logothetis, N.K., and Eschenko, O.
(2018). Locus coeruleus phasic discharge is essential for stimulus-induced
Jodo, E., Chiang, C., and Aston-Jones, G. (1998). Potent excitatory influence
gamma oscillations in the prefrontal cortex. J. Neurophysiol. 119, 904–920.
of prefrontal cortex activity on noradrenergic locus coeruleus neurons.
Neuroscience 83, 63–79. Okun, M., Steinmetz, N., Cossell, L., Iacaruso, M.F., Ko, H., Barthó, P., Moore,
T., Hofer, S.B., Mrsic-Flogel, T.D., Carandini, M., and Harris, K.D. (2015).
Kalwani, R.M., Joshi, S., and Gold, J.I. (2014). Phasic activation of individual
Diverse coupling of neurons to populations in sensory cortex. Nature 521,
neurons in the locus ceruleus/subceruleus complex of monkeys reflects
511–515.
rewarded decisions to go but not stop. J. Neurosci. 34, 13656–13669.
Perkel, D.H., Gerstein, G.L., and Moore, G.P. (1967). Neuronal spike trains and
Kebschull, J.M., Garcia da Silva, P., Reid, A.P., Peikon, I.D., Albeanu, D.F., and stochastic point processes. II. Simultaneous spike trains. Biophys. J. 7,
Zador, A.M. (2016). High-throughput mapping of single-neuron projections by 419–440.
sequencing of barcoded RNA. Neuron 91, 975–987.
Petersen, C.C.H., Hahn, T.T.G., Mehta, M., Grinvald, A., and Sakmann, B.
Lakatos, P., Shah, A.S., Knuth, K.H., Ulbert, I., Karmos, G., and Schroeder, (2003). Interaction of sensory responses with spontaneous depolarization in
C.E. (2005). An oscillatory hierarchy controlling neuronal excitability and stim- layer 2/3 barrel cortex. Proc. Natl. Acad. Sci. USA 100, 13638–13643.
ulus processing in the auditory cortex. J. Neurophysiol. 94, 1904–1911.
Quiroga, R.Q., Nadasdy, Z., and Ben-Shaul, Y. (2004). Unsupervised spike
Lee, S.-H., and Dan, Y. (2012). Neuromodulation of brain states. Neuron 76, detection and sorting with wavelets and superparamagnetic clustering.
209–222. Neural Comput. 16, 1661–1687.
Lee, A., Rosin, D.L., and Van Bockstaele, E.J. (1998). alpha2A-adrenergic Rancic, V., Rawal, B., Panaitescu, B., Ruangkittisakul, A., and Ballanyi, K.
receptors in the rat nucleus locus coeruleus: subcellular localization in cate- (2018). Suction electrode recording in locus coeruleus of newborn rat brain
cholaminergic dendrites, astrocytes, and presynaptic axon terminals. Brain slices reveals network bursting comprising summated non-synchronous
Res. 795, 157–169. spiking. Neurosci. Lett. 671, 103–107.

Neuron 99, 1–14, September 5, 2018 13


Please cite this article in press as: Totah et al., The Locus Coeruleus Is a Complex and Differentiated Neuromodulatory System, Neuron (2018), https://
doi.org/10.1016/j.neuron.2018.07.037

Rash, J.E., Olson, C.O., Davidson, K.G.V., Yasumura, T., Kamasawa, N., and Steriade, M., Nuñez, A., and Amzica, F. (1993). A novel slow (< 1 Hz) oscillation
Nagy, J.I. (2007). Identification of connexin36 in gap junctions between of neocortical neurons in vivo: depolarizing and hyperpolarizing components.
neurons in rodent locus coeruleus. Neuroscience 147, 938–956. J. Neurosci. 13, 3252–3265.
Room, P., Postema, F., and Korf, J. (1981). Divergent axon collaterals of rat Takeuchi, T., Duszkiewicz, A.J., Sonneborn, A., Spooner, P.A., Yamasaki, M.,
locus coeruleus neurons: demonstration by a fluorescent double labeling tech- Watanabe, M., Smith, C.C., Fernández, G., Deisseroth, K., Greene, R.W., and
nique. Brain Res. 221, 219–230. Morris, R.G.M. (2016). Locus coeruleus and dopaminergic consolidation of
Rubinov, M., and Sporns, O. (2010). Complex network measures of brain con- everyday memory. Nature 537, 357–362.
nectivity: uses and interpretations. Neuroimage 52, 1059–1069. Trenholm, S., McLaughlin, A.J., Schwab, D.J., Turner, M.H., Smith, R.G.,
Safaai, H., Neves, R., Eschenko, O., Logothetis, N.K., and Panzeri, S. (2015). Rieke, F., and Awatramani, G.B. (2014). Nonlinear dendritic integration of elec-
Modeling the effect of locus coeruleus firing on cortical state dynamics and trical and chemical synaptic inputs drives fine-scale correlations. Nat.
single-trial sensory processing. Proc. Natl. Acad. Sci. USA 112, 12834–12839. Neurosci. 17, 1759–1766.
Sakata, S., and Harris, K.D. (2009). Laminar structure of spontaneous and sen- Uematsu, A., Tan, B.Z., Ycu, E.A., Cuevas, J.S., Koivumaa, J., Junyent, F.,
sory-evoked population activity in auditory cortex. Neuron 64, 404–418. Kremer, E.J., Witten, I.B., Deisseroth, K., and Johansen, J.P. (2017).
Sakata, S., and Harris, K.D. (2012). Laminar-dependent effects of cortical state Modular organization of the brainstem noradrenaline system coordinates
on auditory cortical spontaneous activity. Front. Neural Circuits 6, 109. opposing learning states. Nat. Neurosci. 20, 1602–1611.
Sánchez-Meca, J., Marı́n-Martı́nez, F., and Chacón-Moscoso, S. (2003). Usher, M., Cohen, J.D., Servan-Schreiber, D., Rajkowski, J., and Aston-Jones,
Effect-size indices for dichotomized outcomes in meta-analysis. Psychol. G. (1999). The role of locus coeruleus in the regulation of cognitive perfor-
Methods 8, 448–467. mance. Science 283, 549–554.
Sara, S.J., and Bouret, S. (2012). Orienting and reorienting: the locus coeruleus Valentino, R.J., and Foote, S.L. (1987). Corticotropin-releasing factor disrupts
mediates cognition through arousal. Neuron 76, 130–141. sensory responses of brain noradrenergic neurons. Neuroendocrinology
Sara, S.J., and Segal, M. (1991). Plasticity of sensory responses of locus 45, 28–36.
coeruleus neurons in the behaving rat: implications for cognition. Prog. Brain Van Bockstaele, E.J., Garcia-Hernandez, F., Fox, K., Alvarez, V.A., and
Res. 88, 571–585. Williams, J.T. (2004). Expression of connexins during development and
Sceniak, M.P., and Maciver, M.B. (2006). Cellular actions of urethane on rat following manipulation of afferent input in the rat locus coeruleus.
visual cortical neurons in vitro. J. Neurophysiol. 95, 3865–3874. Neurochem. Int. 45, 421–428.
Schwarz, L.A., Miyamichi, K., Gao, X.J., Beier, K.T., Weissbourd, B., DeLoach, van den Pol, A.N., Ghosh, P.K., Liu, R.-J., Li, Y., Aghajanian, G.K., and Gao,
K.E., Ren, J., Ibanes, S., Malenka, R.C., Kremer, E.J., and Luo, L. (2015). Viral- X.-B. (2002). Hypocretin (orexin) enhances neuron activity and cell synchrony
genetic tracing of the input-output organization of a central noradrenaline in developing mouse GFP-expressing locus coeruleus. J. Physiol. 541,
circuit. Nature 524, 88–92. 169–185.
Shipley, M.T., Fu, L., Ennis, M., Liu, W.L., and Aston-Jones, G. (1996). van Welie, I., Roth, A., Ho, S.S.N., Komai, S., and Ha €usser, M. (2016).
Dendrites of locus coeruleus neurons extend preferentially into two pericoer- Conditional spike transmission mediated by electrical coupling ensures milli-
ulear zones. J. Comp. Neurol. 365, 56–68. second precision-correlated activity among interneurons in vivo. Neuron 90,
Simpson, K.L., Altman, D.W., Wang, L., Kirifides, M.L., Lin, R.C., and 810–823.
Waterhouse, B.D. (1997). Lateralization and functional organization of the Vankov, A., Hervé-Minvielle, A., and Sara, S.J. (1995). Response to novelty and
locus coeruleus projection to the trigeminal somatosensory pathway in rat. its rapid habituation in locus coeruleus neurons of the freely exploring rat. Eur.
J. Comp. Neurol. 385, 135–147. J. Neurosci. 7, 1180–1187.
Smeets, W.J., and González, A. (2000). Catecholamine systems in the brain of Wagatsuma, A., Okuyama, T., Sun, C., Smith, L.M., Abe, K., and Tonegawa, S.
vertebrates: new perspectives through a comparative approach. Brain Res. (2018). Locus coeruleus input to hippocampal CA3 drives single-trial learning
Brain Res. Rev. 33, 308–379. of a novel context. Proc. Natl. Acad. Sci. USA 115, E310–E316.

14 Neuron 99, 1–14, September 5, 2018


Please cite this article in press as: Totah et al., The Locus Coeruleus Is a Complex and Differentiated Neuromodulatory System, Neuron (2018), https://
doi.org/10.1016/j.neuron.2018.07.037

STAR+METHODS

KEY RESOURCES TABLE

REAGENT or RESOURCE SOURCE IDENTIFIER


Antibodies
Monoclonal mouse anti-tyrosine hydroxylase ImmunoStar RRID: AB_572268
Rat absorbed anti-mouse secondary antibody Biozol RRID: AB_2336180
Chemicals, Peptides, and Recombinant Proteins
Urethane Sigma-Aldrich U2500
Clonidine (0.5 mg/kg intra-peritoneal) Sigma-Aldrich C7897
DAB and horseradish peroxidase reaction with Biozol Vectastain Elite ABC-Peroxidase Kit CE
hydrogen peroxide kit
Slide clearance Carl Roth Roti-histol
Slide mounting media DPX Sigma-Aldrich 06522
Experimental Models: Organisms/Strains
Rats (Sprague-Dawley, male, 350 – 450 g body Charles River Laboratories Sulzfeld, Germany
weight, specific pathogen free)
Software and Algorithms
MATLAB The MathWorks Version 2015a
Digital signal storage and manual spike CED Spike2
sorting curation
Other
Tungesten electrode for microstimulation FHC UEWMFGSMCNNG
Silicone probe for LC recording NeuroNexus A1x32-Poly3-10mm-25 s-177-A32
Amplifier and hardware filter AlphaOmega MPC Plus
Digital signal acquisition CED Power1401mkII

CONTACT FOR REAGENT AND RESOURCE SHARING

Further information and requests for resources and reagents should be directed to and will be fulfilled by the Lead Contact, Oxana
Eschenko (oxana.eschenko@tuebingen.mpg.de).

EXPERIMENTAL MODEL AND SUBJECT DETAILS

Subjects
All experimental procedures were carried out with approval from the local authorities and in compliance with the German Law for the
Protection of Animals in experimental research (Tierschutzversuchstierverordnung) and the European Community Guidelines for the
Care and Use of Laboratory Animals (EU Directive 2010/63/EU). Twelve male Sprague-Dawley rats (350 - 450 g) were used. Animals
were ordered from Charles River Laboratories (Sulzfeld, Germany) as specific pathogen free rats. Animals were pair housed in IVC
and on a 08:00 to 20:00 dark to light cycle.

METHOD DETAILS

Anesthesia and Surgical Procedures


Rats were anesthetized using an intra-peritoneal (i.p.) injection of a 1.5 g/kg body weight dose of urethane (Sigma-Aldrich, U2500).
Oxygen was administered. The animal was placed on a heating pad and a rectal probe was used to maintain a body temperature of 37
C. The eyes were covered in ointment. After removal of the skin, the skull was leveled to 0 degrees, such that the difference between
lambda and bregma was less than 0.2 mm.

Neuron 99, 1–14.e1–e6, September 5, 2018 e1


Please cite this article in press as: Totah et al., The Locus Coeruleus Is a Complex and Differentiated Neuromodulatory System, Neuron (2018), https://
doi.org/10.1016/j.neuron.2018.07.037

Stereotaxic coordinates and electrode placement


Craniotomies were made at the locations listed in the following table (Table S1). Accurate electrode placement was confirmed by
examining the firing properties of neurons in each brain region. In the LC, these criteria included a slow spontaneous firing rate,
biphasic response to noxious sensory stimuli (foot shock), audible presence of jaw movement-responsive cells in the MeV
(Mesencephalic Nucleus of Cranial Nerve V) with undetectable single units (< 0.2 mV). LC electrode placements were later verified
using histological examination of brain tissue sections (Figure S9). Placement of electrodes for forebrain stimulation was stereotacti-
cally-guided and, when possible, electrophysiological criteria were used to verify target placement.

Electrodes
Stimulation of cortical and sub-cortical brain regions was conducted via tungsten electrodes with low impedance (10 - 50 kOhm) to
prevent heating at the electrode tip. Tungsten electrodes with a diameter of 200 mm (FHC, Model: UEWMFGSMCNNG) were ground
at the tip to lower impedance to this range. Recording from the LC used a 15 mm thick silicone probe with 32 channels (NeuroNexus,
Model: A1x32-Poly3-10mm-25 s-177-A32). The channels were implanted toward the anterior aspect of the brain. Each channel was
separated from the neighboring channels by 25 mm. Channels were divided into 10 tetrodes with one channel overlapping per tetrode
(Figure S1). When placing our arrays, we advanced ventrally until after the biphasic response (excitation, followed by self/lateral
inhibition) to sensory stimuli occurred clearly on all channels. According to this criteria, we found that it was possible to have all
electrodes (spanning 275 mm in dorso-ventral axis) within the LC core, which spans 500 mm (Grzanna and Molliver, 1980; Shipley
et al., 1996). Our array was therefore advanced to the interior of the LC and could not span the entire extent of the LC.

Recording and signal acquisition


A silver wire inserted into the neck muscle was used as a reference for the electrodes. Electrodes were connected to a pre-amplifier
(in-house constructed) via low noise cables. Analog signals were amplified (by 2000 for LC and 500 for cortex) and filtered (8 kHz low
pass, DC high pass) using an Alpha-Omega multi-channel processor (Alpha-Omega, Model: MPC Plus). Signals were then digitized
at 24 kHz using a data acquisition device CED, Model: Power1401mkII). These signals were stored using Spike2 software (CED).

Characterization of LC units
Single units were identified using typical criteria (Figure S1). These criteria were low firing rate (0.89 spikes/sec) and a characteristic
bi-phasic response (excitation followed by inhibition) to sensory stimuli. The neurochemical nature of LC units was identified by the
presence of auto-inhibitory alpha-2 adrenergic receptors, which were activated using the alpha-2 agonist, clonidine. Electrode tracks
approaching LC were visualized (Figure S9). These methods are the standard for demonstrating that the neurons are likely to be
LC-NE neurons. Therefore, our results are comparable with the existing literature on the LC and we were likely sampling mostly
LC-NE neurons.

Sensory stimuli
Sensory stimuli were foot shocks delivered to the contralateral hind paw. Pulses were square, biphasic, and 0.5 msec duration at
5 mA. Pulses were delivered at two frequencies (single pulse or five pulses at 30 Hz) delivered in random order. Fifty trials of foot
shock stimuli were delivered with an inter-trial interval of 2000 ± 500 msec.

Intra-cranial stimulation
Brain regions were stimulated in a random order. Pulses were square, biphasic, and 0.25 msec duration over a range of intensities
(400 - 1200 mA), which were delivered in a randomized order. The pulse waveforms were constructed in Spike2 (CED) and delivered
via a current generator (in-house constructed), which allowed recording of the stimulation voltage at the tip of electrode, which was
also digitized and stored and used to verify stimulation. Stimulation was delivered with an inter-trial interval of 2000 ± 500 msec. At
least 50 trials of each intensity were delivered for each brain region.

Administration of clonidine
At the end of the recording, a 0.05 mg/kg dose of the alpha-2 adrenergic agonist clonidine was injected i.p. (Sigma-Aldrich, product
identification: C7897). The recording was continued at least until all activity, included multi-unit activity, ceased. Clonidine adminis-
tration allows unambiguous discrimination of LC neurons from surrounding structures within the recording range of our electrodes
because they do not have alpha-2 receptors (McCune et al., 1993).

Histology
Rats were euthanized with pentobarbital sodium (Narcoren, Merial) via an i.p. injection (100 mg/kg). The rats were then trans-cardially
perfused with 0.9% saline and then 4% paraformaldehyde (PAF) in 0.1M phosphate buffer (pH 7.4). The brain was removed and
stored in 4% PAF. Brains were moved into 30% sucrose, until they sank, before sectioning on a freezing microtome (Microm, model:
HM 440E). Coronal sections (50 mm thick) were collected into 0.1M phosphate buffer (PB) and directly stained. For sections contain-
ing the LC, alternating sections were stained for Nissl substance or the catecholamine synthesis enzyme, tyrosine hydroxylase (TH).
Sections containing cortical and sub-cortical regions were stained for Nissl substance. Staining for TH was performed using a 1:4000

e2 Neuron 99, 1–14.e1–e6, September 5, 2018


Please cite this article in press as: Totah et al., The Locus Coeruleus Is a Complex and Differentiated Neuromodulatory System, Neuron (2018), https://
doi.org/10.1016/j.neuron.2018.07.037

dilution of monoclonal mouse anti-TH antibody (ImmunoStar, RRID: AB_572268) and a 1:400 dilution of biotinylated, rat absorbed
anti-mouse secondary antibody (Biozol, RRID: AB_2336180) in PB. The antibody was visualized using a DAB and horse radish perox-
idase reaction with hydrogen peroxide using a standard kit (Biozol, model: Vectastain Elite ABC-Peroxidase Kit CE). After staining for
TH or Nissl, sections were mounted on gelatin-coated glass slides. Nissl stained and slide mounted sections were dehydrated in an
alcohol series. Slides were cleared (Carl Roth, Roti-histol) and coverslipped with DPX slide mounting media (Sigma-Aldrich, catalog
number: 06522).

QUANTIFICATION AND STATISTICAL ANALYSIS

The statistical tests used and the resulting test statistics and p values, as well as power calculations, may be found in the main text of
the results section. Data were tested for normality using a Shapiro-Wilk test (alpha = 0.05) and homogeneity of variance (alpha = 0.05)
using an F-test (vartest2 in MATLAB) for 2 groups or Levene’s Test (vartestn in MATLAB) for more than 2 groups. If data were not
normal, then a Wilcoxon-Mann-Whitney Test was used for 2 groups or a Kruskal-Wallis Test for more than 2 groups; otherwise, a
two-sided t test or between-subjects one-way ANOVA was used. If variance was inhomogeneous, then we used Welch’s t test or
Welch’s ANOVA. For ANOVA (or Kruskal-Wallis), post hoc testing between individual groups was performed using the MATLAB
function, multcompare (with Dunn-Sidak correction for Kruskal-Wallis). If a Welch’s ANOVA was used for heteroscedastic data,
then post hoc testing was performed using the Games-Howell test. These comparisons were unplanned. Mean and standard error
are reported for normally distributed data. Median and standard deviation are reported for data that were not normally distributed.
We report effect sizes as Cohen’s D for analysis of 2 groups (e.g., t test or Wilcoxon-Mann-Whitney) or, for analysis of more than 2
groups, we report u2 (e.g., ANOVA) or u2adj (e.g., Welch’s ANOVA). For 2 3 2 tables, we used a Fisher’s Exact Test, for which the
effect size is quantified by the Odds Ratio (OR), which we converted to Cohen’s D (termed Dor) (Sánchez-Meca et al., 2003).
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
m  m2 ðn1  1Þs21 + ðn2  1Þs22
D= 1 ; Sp =
Sp n1 + n2  2

pffiffiffi
3
Dor = lnðORÞ 3
p

 
SSbetween groups  dfbetween groups 3 Mean Squared Error
u2 =
Mean Squared Error + SStotal

dfbetween groups 3 ðF  1Þ
u2adj = ; N = number of samples
dfbetween groups 3 ðF  1Þ + N

In the case that a null hypothesis was rejected, we made a post hoc determination that the sample size and statistical test provided
adequate power to reject the null hypothesis. The power was calculated with sampsizepwr in MATLAB for 2 groups. If there were
more than 3 groups, then the power was calculated using powerAOVI in MATLAB. We are unaware of methods for assessing the
power of a Kruskal-Wallis Test or a Welch’s ANOVA and do not report power for those tests.
For circular data, uniformity was assessed using Rayleigh’s Test for Circular Uniformity (alpha = 0.05). These calculations were
made using the CircStat toolbox in MATLAB (Berens, 2009). Power calculations were not made for circular statistics.
The reported sample sizes (N) refer to either rats or single units or single unit pairs and is specified with each usage. Standard errors
and means are reported, unless otherwise noted. Boxplots (in Figures S2, S3, and S7) illustrate the median (dot), the 25th and 75th
percentiles (box), the most extreme points (whiskers), and outliers (open circles).

Spike detection
The recorded signal for each channel was filtered offline with a four pole butterworth band pass (300 - 8000 Hz). Spikes were then
detected as crossings of a negative threshold that was four times the standard deviation of the channel noise. Noise was defined as
the median of the rectified signal divided by 0.6745 (Quiroga et al., 2004). Detected spike waveforms were stored from 0.6 msec to
1.0 msec around the threshold crossing. This duration was chosen based on the known action potential duration of rat LC neurons
(Aston-Jones and Bloom, 1981a; Sara and Segal, 1991). A 0.6 msec refractory period was used to not detect a subsequent spike
during this window.

Neuron 99, 1–14.e1–e6, September 5, 2018 e3


Please cite this article in press as: Totah et al., The Locus Coeruleus Is a Complex and Differentiated Neuromodulatory System, Neuron (2018), https://
doi.org/10.1016/j.neuron.2018.07.037

Spike clustering
Spike waveforms were clustered using an automatic clustering algorithm and then manually refined and verified using cluster visu-
alization software (CED, Spike 2). Automated clustering was performed using Wave_Clus (Quiroga et al., 2004) in MATLAB (default
parameters for clustering) followed by manual refinement and verification in clustering visualization software. This method uses
wavelets to decompose the waveform into a simpler approximation of its shape at different frequencies (wavelet scales) and times
in the waveform. Using this method, small amplitude bumps or deflections at different time points in the waveform can be used to
cluster waveforms together, if they are a highly informative waveform characteristic. After the automated sorting, manual refinement
of clustering using a 3-dimensional plot of principle components or the amplitude at particular waveform time points (peaks and
troughs). Auto-correlograms were used to assess the level of noise (refractory period violations) and cross-correlograms between
simultaneously recorded units were used to prevent over-clustering.

Detection of spikes across tetrodes


Due to configuration of the recording array with a high-density of electrode contacts, the spikes from the same LC neuron could be
detected on more than one tetrode. In such situations, we first attempted to merge the spike trains across multiple tetrodes. Merging
the spike trains potentially originating from the same neuron and detected on multiple tetrodes should reduce false negatives
(missing spikes), as it is common for PCA of waveforms to miss some spikes even if they are part of a well isolated cluster. The
assumption is that different spikes are missed on different tetrodes, which were subjected to separate PCA’s during clustering.
Therefore, spikes missed by PCA on one tetrode could be partly filled in by spikes that were detected on other tetrodes, providing
that the tetrodes were recording the same single unit. Merging spike trains across tetrodes operated on the principle that, if the spikes
from the same neuron are recorded, for example, on 3 adjacent tetrodes and the spike waveforms can be classified into 3 well-iso-
lated clusters, then the spike trains can be merged to yield an equally well-isolated unit. Furthermore, merging across spike trains
allows units to be tracked if they drift away from one tetrode and become closer to another tetrode. However, the merging procedure
needs to avoid inclusion of contaminated spikes originating from neighboring cells. Therefore, merged spike trains must be statis-
tically tested for false positives and conservatively discarded. Unit cluster contamination from other units are typically detected
by the presence of spikes during the refractory period. Thus, if the merged spike train did not meet criteria for a single unit activity,
then we kept the unit recorded on the tetrode with the least noise (lowest proportion of spikes during refractory period).
The merging process consisted in the following steps. First, the cross-correlograms were computed at the sampling rate of the
recording (0.04 ms bin width) between the spike trains of a cluster isolated from one tetrode (‘‘reference’’ cluster) and all other clusters
isolated from all remaining tetrodes. If the spike trains associated with the two clusters contained spikes from the same unit, then the
majority of spikes would have identical timing with the vast majority (> 90%) of spikes being coincidental at time 0 (with a few sampling
points of error) and the remaining spikes being spikes either detected on one tetrode and missed on the other or cluster noise from
other units. Prior recordings using high-density linear electrodes have used cross-correlograms to simply discard one of the trains;
however, we used merging across tetrodes to reduce missed spikes. In the case that the ‘‘reference’’ and ‘‘other’’ spike trains were
mostly coincidental spikes, we attempted to merge them using a procedure, as follows. The coincident spikes in the reference spike
train were deleted and the remaining reference spikes were merged with the spikes from the other spike train, resulting in a new
‘‘merged’’ spike train. The amount of noise (number of spikes during the refractory period) and total number of spikes in the train
were recorded for the original ‘‘reference’’ spike train, original ‘‘other’’ spike train, and newly merged spike train. A Fisher’s Exact
Test was performed to statistically assess if the proportion of contaminated spikes added by merging is significantly greater than
the proportion of noise (refractory period) spikes in either the original reference spike train or the original other spike train. The Fisher’s
test was run separately for merged versus reference and the merged versus other spike trains. A result of non-significance (with alpha
set to 0.01 and a right-sided test) for both the original reference spike train and the original other spike train indicated that the original
spike trains do not have greater odds of having noise than the merged spike train. In this case, the merged train may be kept because
the amount of noise was not increased beyond its level in the original two trains. The original spike trains were discarded. However, if
the test is significant for either of the original trains, then either the reference spike train or the other spike train is kept depending on
which has a lower percentage of spikes during the refractory period out of the total number of spikes. The merged spike train was
discarded. The merging procedure was repeated for each cluster from each tetrode until conflicts no longer existed.

Assigning unit location on recording array


Isolated single units were assigned a channel location on the electrode array according to which electrode measured the highest
mean waveform amplitude (averaged from all spikes). In the case of single units with spikes that were merged across tetrodes, a
list tracked the tetrodes on which the unit appeared and the location was assigned to the channel with the maximum amplitude
when considering all of the tetrodes that recorded the unit. The spacing between all 32 channels was 25 mm, which allowed us to
use Pythagoras’ Theorem to calculate a distance between channels on the array. The distance between each unit’s maximal wave-
form amplitude was used to measure the distance dependence of spike count correlations and cross-correlograms.
We inferred the spatial probability distribution of narrow and wide units on the array by fitting Nth-order polynomials (n = 2 to 9) to
the proportion of each unit type recorded at each of the 10 tetrodes. We found that n = 5 or 6 produced an R2 that was > 0.9, whereas

e4 Neuron 99, 1–14.e1–e6, September 5, 2018


Please cite this article in press as: Totah et al., The Locus Coeruleus Is a Complex and Differentiated Neuromodulatory System, Neuron (2018), https://
doi.org/10.1016/j.neuron.2018.07.037

lower N had poor fits (R2 = 0.3 to 0.7) and larger N visually overfit the data. The polynomial function provided a probability, which we
verified in all rats by removing one rat and re-calculating the fit until all rats had been removed once (jackknife error). In all cases,
except one, R2 was > 0.9.

Spike count correlations


The Pearson’s correlation coefficient was used to quantify the correlation between spike counts (Figures 2, 7, S2, and S3). Sponta-
neous spiking excluded the 1 s period following foot shock stimulation or intra-cranial stimulation. Spontaneous spikes were also
excluded during inactive periods in which the rate was less than 0.5 Hz due to quiescence of all single and multi-unit activity in
the LC (paradoxical sleep; Aston-Jones and Bloom, 1981a). Spontaneous spike count correlations were then calculated from the
time bins (200 msec or 1000 msec) in which both neurons were active.
Poisson spike trains should generate some degree of synchrony (spike count correlation coefficient) by chance. We compared the
correlation coefficient for each pair against 500 surrogate spike trains for the same pair. The trains were generated in manner that
preserved the inter-spike interval (ISI) distribution for each unit and the slow (< 2 Hz) oscillations in spike rate (thought to be generated
by brain network-wide interactions under anesthesia). Each surrogate spike train had the same oscillation phase preference, the
same number of spikes per oscillation cycle, and an identical ISI distribution, but the precise spike timing of each unit in relation
to other units is randomized.
Evoked spike count correlations after foot shocks were calculated from the trial-by-trial spike count in a single window after
stimulus onset (50 msec). This window was chosen based upon the timing of the spiking evoked by a single pulse foot shock in
our recordings, as well as reports by others (Aston-Jones and Bloom, 1981b; Foote et al., 1980). For evoked correlations, the spiking
on each trial was sparse in the 50 msec window, so a statistical permutation test was under-powered and could not be performed.

Cross-correlograms
We calculated cross-correlograms between spike trains (Figures 4 and S5). Significant changes in coincidental spike count were
detected by comparing the observed counts to 1000 surrogate cross-correlograms generated from jitter spike times (Fujisawa
et al., 2008). This approach uses the data to determine the degree of coincident spiking expected by chance and it also excludes
synchrony due to interactions at slower timescales than those of interest. Briefly, the spike times for each unit were jittered on a
uniform interval and a surrogate cross-correlogram was calculated between the jittered spike times; this process was repeated
1000 times. Significant cross-correlogram bins were those that crossed both a 1% pairwise expected spike count band and a
1% globally expected spike count band (the maximum and minimum counts at any time point in the cross-correlogram and any
surrogate cross-correlogram). For broad-type interactions, the cross-correlograms were calculated in a window of 2000 msec, a
bin size of 10 msec, and a uniform jitter interval of ± 200 msec. Any significant coincidental spiking excludes synchrony due to
co-variation of spiking on a timescale of a few hundred milliseconds. For sharp-type interactions, we used a window of 3 msec, a
bin size of 0.05 msec, and a uniform jitter interval of ± 1 msec.
To ascertain if we missed interactions at other timescales, we employed the method of integrating the cross-correlogram (Bair
et al., 2001). The gradual integration of the cross-correlogram in successively larger windows (e.g., t = ± 5 msec,
10 msec, .40,000 msec) will result in a curve that changes rapidly during tau with a large concentration of coincidental spiking
and will eventually plateaus at very large tau at which the firing patterns of the pair is unrelated. We integrated in 1 msec steps
from 0 to 40 s and recorded the start of the plateau. The results of this analysis suggested that some interactions could occur on
the level of a few seconds. Thus, we calculated additional cross-correlograms using a window of 20 s, a bin size of 0.2 s, and a
uniform jitter on the interval of ± 2 s.

Syn-fire chain analysis


Repeating sequences of triplets of neuronal spiking were measured using three steps: (i) a spike-by-spike search, (ii) template-
formation, and (iii) template matching algorithm (Ikegaya et al., 2004; Luongo et al., 2016). The analysis stepped through all simulta-
neously recorded units n/N and all of the spike times (M) for each unit (nm/M). The spike search started with the first unit and its first
spike, nm . This spike time was a reference event marking the start of a 2 msec window during which we identified any other spiking
units. The sequence of units and the delay between their spikes was stored as a template. For example, nm might be followed
1.1 msec later by a spike from unit n = 5 which is subsequently followed 0.8 msec later by a spike from unit n = 30. This forms a
template of 1 – 5 – 30 with delays of 1.1 msec and 0.8 msec. The next step, template matching, would proceed through the spikes
1m+1/M and attempt to match the template and its delays. A 0.4 msec window of error was allowed around each spike in the original
template for matching. Thus, for each spike of unit n = 1, unit n = 5 could spike between 0.7 msec and 1.5 msec after unit n = 1 and
unit n = 30 could spike 0.4 msec to 1.2 msec after unit n = 5. A template match would be counted if the spikes of the other units aligned
with the originally formed template. For each template, the total number of observations was compared to the number of observa-
tions in the top 1% of 100 surrogate datasets in which spike times were jittered on a uniform interval by 1 msec. Any sequential spike
patterns that occurred more often than expected by chance were counted as significantly occurring chains of spikes.

Neuron 99, 1–14.e1–e6, September 5, 2018 e5


Please cite this article in press as: Totah et al., The Locus Coeruleus Is a Complex and Differentiated Neuromodulatory System, Neuron (2018), https://
doi.org/10.1016/j.neuron.2018.07.037

Population coupling
We followed the method of Okun et al. (2015) for assessing synchrony between individual single units and the surrounding population.
After selecting one single unit, we merged the spike trains of the remaining single units into a population spike train. We then
computed the cross-correlogram between the selected single unit and the population. The cross-correlogram was calculated
with 1 msec bins over a period of ± 400 msec and Z-score normalized to the period between 300 and 400 msec (edges of the cross-
correlogram). A large Z-score at time 0 would indicate coupling of the single unit to the population. A Z-score of 0 would reflect no
relation to the population (Figure 6A, red line for unit B) and would indicate ‘‘uncoupling.’’ A large, positive Z-score would reflect that
each time a single unit spikes, many other simultaneously recorded neurons also spike (Figure 6A, blue line for unit A). A large,
negative Z-score would reflect less spiking of the population around the time a single unit spikes. Thus, negative Z-scores are not
uncoupling, rather they indicate inhibitory coupling.

Graph theory analysis and ensemble detection


This analysis was used for Figures 6 and S7. For each rat, a graph was constructed with each unit as a node. Links were drawn
between units with strong spike count correlations (Rubinov and Sporns, 2010). The threshold for drawing a link was set as the
highest possible value such that the mean network degree was less than the log(N), where N is the number of nodes in the graph.
Units without strongly correlated activity were left unlinked. The resulting network was represented by a binary adjacency matrix.
Ensembles were detected by segregating nodes into groups that maximize the number of links within each group and minimizes
the number of links between group. The iterative optimization procedure used a Louvain community detection algorithm to maximize
a modularity score (Q) quantifying the degree to which groups separate. The degree of ensemble separation (Q) was compared to
modularity scores from 1000 from shuffled networks. If Q was higher than the top 5% of the 1000 surrogate values, then adequate
separation of units into ensembles was achieved. All graph theory and ensemble detection analyses were implemented in MATLAB
using the Brain Connectivity Toolbox (Bullmore and Sporns, 2009).

Oscillations in spike count


Single unit spike trains were first convolved with a Gaussian kernel with a width of 250 msec and a sampling rate of 1 msec. The
resulting spike density functions (SDF) were analyzed for the power of oscillations, the phase of oscillations, and the coherence
between pairs of SDF’s. The power spectral density of each spike train was calculated using a multi-taper method in MATLAB
(Chronux Toolbox) (Mitra and Bokil, 2007). We used 19 tapers with a time-bandwidth product of 10. The frequency band was
0.01 to 10 Hz. We used finite size correction for low spike counts. Instantaneous phase was extracted by first filtering the SDF
with a 3rd order Butterworth filter at a particular frequency of interest (0.09 to 0.11 Hz and 0.40 to 0.48 Hz) and then obtaining phase
from the Hilbert transform of the signal. The consistency of the instantaneous phase difference between each unit pair was assessed
using Rayleigh’s Test for Circular Uniformity (p < 0.05), which was implemented in MATLAB (CircStat Toolbox) (Berens, 2009). Coher-
ence between pairs of units was calculated using the Chronux Toolbox in MATLAB with the same parameters used for calculating the
power spectral density. Coherence and power were averaged across all single units and smoothed with a width of 0.15 Hz.

Spike-LFP phase locking


This analysis was used in Figure 5. Local field potential (LFP) was recorded in the prefrontal cortex. The LFP was lowpass filtered with
a 3rd order Butterworth filter at 2 Hz. The instantaneous phase was obtained by Hilbert transformation. Spike times corresponded to
LFP phases. For each single unit, the phase distribution was tested for significant locking to cortical LFP using Rayleigh’s Test for
Circular Uniformity (p < 0.05).

Antidromic spiking
This analysis was used in Figures 7 and S8. Forebrain regions were stimulated in a randomized order with single pulses at currents of
400, 600, 800, 1000, and 1200 mA. Stimulation was delivered with a 2 s inter-trial interval with 500 msec jitter. Peri-stimulus time
histograms were Z-scored to 1 s before stimulus onset. If a single bin (5 msec), but no other bins, exceeded a Z-score of 5, then
the unit was marked as antidromically activated. The bin size was chosen, based on prior work, which has demonstrated that 3
or 4 msec of jitter occurs when stimulating LC fibers because they lack myelin (Aston-Jones et al., 1985). This rationale is based
on an extremely low chance of consistent spiking within the same 5 msec window by slowly firing neurons (typical ISI is longer
than 100 msec). However, manual inspection of the individual spike rasters was used to confirm the results.

DATA AND SOFTWARE AVAILABILITY

The generated datasets and MATLAB codes for analysis are available from the Lead Contact on reasonable request.

e6 Neuron 99, 1–14.e1–e6, September 5, 2018

Das könnte Ihnen auch gefallen