You are on page 1of 9

Review TRENDS in Neurosciences Vol.30 No.

INMED/TINS special issue

Are corticothalamic ‘up’ states


fragments of wakefulness?
Alain Destexhe1, Stuart W. Hughes2, Michelle Rudolph1 and Vincenzo Crunelli2
1
CNRS, Integrative and Computational Neuroscience Unit 1, Avenue de la Terrasse, Gif sur Yvette Cedex 91198, France
2
School of Biosciences, Cardiff University, Museum Avenue, Cardiff CF10 3US, UK

The slow (<1 Hz) oscillation, with its alternating ‘up’ and The dynamics of single neurons during ‘activated’
‘down’ states in individual neurons, is a defining feature of states and slow-wave sleep
the electroencephalogram (EEG) during slow-wave sleep We start by reviewing the essential cellular correlates of the
(SWS). Although this oscillation is well preserved across ‘activated’ brain state, which corresponds to attentive wake-
mammalian species, its physiological role is unclear. fulness, and examine how they qualitatively compare with
Electrophysiological and computational evidence from those of SWS. Brain activation is invariably associated with
the cortex and thalamus now indicates that slow-oscil- a so-called ‘desynchronized’ electroencephalogram (EEG),
lation ‘up’ states and the ‘activated’ state of wakefulness which consists of low-amplitude fluctuations at relatively
are remarkably similar dynamic entities. This is consistent high frequencies (>15 Hz) [5,6] (Figure 1a). This state is
with behavioural experiments suggesting that slow-oscil- associated with tonic and irregular firing of cortical neurons
lation ‘up’ states provide a context for the replay, and [19], the membrane potential of which has been shown in
possible consolidation, of previous experience. In this naturally waking animals to be persistently depolarized and
scenario, the T-type Ca2+ channel-dependent bursts of to fluctuate around 60 mV [5,6] (Figure 1a).
action potentials that initiate each ‘up’ state in thalamo- During SWS, the overriding activity of the EEG is the
cortical (TC) neurons might function as triggers for synap- so-called slow (<1 Hz) sleep rhythm or slow (<1 Hz) sleep
tic and cellular plasticity in corticothalamic networks. This oscillation [2,3,20] (Figure 1a). This slow oscillation com-
review is part of the INMED/TINS special issue Physio- prises rhythmically repeating, large-amplitude biphasic
genic and pathogenic oscillations: the beauty and the waves (i.e. slow waves) and is responsible for temporally
beast, based on presentations at the annual INMED/TINS
symposium (http://inmednet.com). Glossary

Introduction
In vitro models of the slow oscillation: a slow oscillation can be
The function of sleep is a mystery that has long fascinated reliably generated in isolated slices of the neocortex. This oscil-
biologists and is still the matter of intense debate [1]. One lation is an emergent property of networks of cortical neurons, with
of the most prominent features of sleep in mammals is the the ‘up’ state being generated by recurrent excitation that is
occurrence of the slow (<1 Hz) sleep oscillation that dom- balanced and regulated by inhibitory networks [24]. The ‘down’
state arises if these mechanisms fail but also seems to be shaped by
inates slow-wave sleep (SWS; Box 1). This oscillation is
activity-dependent (i.e. Ca2+, Na+ and ATP-dependent) K+ channels
extremely similar in different species [2–12], suggesting [24,42]. A recovery from this active disfacilitation can eventually
that it might have well-defined, conserved roles. Fairly lead to a new ‘up’ state. In thalamic slices, individual TC and TRN
recently, behavioural experiments have indicated that neurons can generate a slow oscillation intrinsically, without the
SWS might be related to memory consolidation [11–17] need for any network input. This occurs if leak K+ currents are
reduced below a certain threshold [43,44], a feat that can be easily
and the basis of such consolidation might be the slow sleep
achieved by tonically activating metabotropic glutamate receptors,
oscillation itself [12,14–18]. Here, we explore this issue as probably occurs in the whole brain. In both TC and TRN neurons,
from an electrophysiological and computational modelling the ‘up’ and ‘down’ states are primarily generated by the ‘switch-
perspective. Specifically, we reassess various electro- ing’ on and off, respectively, of the T-type Ca2+ ‘window’ current
physiological measurements of the waking (or wake-like) (reviewed in Refs. [9,43–47]).
Local field potentials (LFPs): recordings of extracellular electrical
state and compare them with those obtained for the ‘up’ activity using low-impedance electrodes. They provide a signal
state of the slow (<1 Hz) sleep oscillation in the specific similar to the EEG but sample more local populations (of the order
brain structures involved in its generation. The extremely of 1 mm) in the cortex.
close similarity of both single-neuron and network a rhythms: EEG oscillations at 10 Hz (range, 8–13 Hz) that prim-
dynamics during these different scenarios is compatible arily occur during relaxed wakefulness. In humans, they are most
pronounced at occipital sites and most prominent when the eyes
with the results of behavioural experiments indicating that are closed.
during SWS selected epochs of prior experience are episo- Sleep spindles: brief bursts (typically lasting 0.5–1.5 s) of rhythmic
dically replayed and consolidated. EEG activity at 7–14 Hz. Sleep spindles are characteristic of early
sleep and commonly associated with individual slow-wave com-
Corresponding author: Crunelli, V. (crunelli@cardiff.ac.uk). plexes (see above).
Available online 3 May 2007.

www.sciencedirect.com 0166-2236/$ – see front matter ß 2007 Elsevier Ltd. All rights reserved. doi:10.1016/j.tins.2007.04.006
Review TRENDS in Neurosciences Vol.30 No.7 335

Box 1. The EEG slow (<1 Hz) oscillation and its relationship to neuronal activity and other types of sleep rhythms (Figure I).

Figure I. (a) One cycle of the EEG slow (<1 Hz) oscillation at different cortical depths. Note the reversal in polarity between 0.25 and 0.5 mm. The bottom panel shows an
enlarged slow wave from a depth recording with depth-positive and -negative components clearly indicated. Modified and reproduced, with permission, from [61]. (b) A
simultaneous surface EEG, depth EEG and intracellular recording from a cortical pyramidal neuron during the slow oscillation. The depth-positive EEG component is
associated with neuronal silence (‘down’ state), whereas the depth-negative component is related to neuronal depolarization (‘up’ state). Modified and reproduced, with
permission, from [33]. (c,d) TC and TRN neurons also show ‘up’ and ‘down’ states in association with the EEG depth-negative and -positive components, respectively.
Note the stereotypical nature of the ‘up’ and ‘down’ states and how each ‘up’ state commences with a T-type Ca2+ channel-mediated burst. Sleep spindles tend to be
grouped by the slow oscillation and typically occur immediately following the depth-negative peak (d). They are closely associated with rhythmic firing during the early
part of the ‘up’ state in TRN neurons. Although sleep spindles are primarily generated within the TRN, their widespread spatiotemporal coherence is brought about by
corticothalamic feedback [65]. Modified and reproduced, with permission, from [33,66]. (e) Bursts of EEG d (1–4 Hz) rhythms recur within the frequency of the slow
oscillation. These rhythms can be traced back to the activity of several TC neurons, which can exhibit brief epochs of intrinsic rhythmic bursting rather than a silent
‘down’ state. Again, although d-frequency activities originate primarily from intrinsic thalamic oscillators, these oscillators are synchronized by corticothalamic feedback
[29]. In (c) and (d), EEG and intracellular recording were performed simultaneously, whereas the upper and lower traces in (e) are from separate recordings. Modified
and reproduced, with permission, from [21,22]. All data are from the cat.

grouping all other types of sleep-related EEG patterns, correlated ‘up’ and ‘down’ states in individual cortical
such as spindle waves (7–14 Hz) and d oscillations neurons that are similar to those that occur during natural
(1–4 Hz), into clearly demarcated, recurring sequences sleep [10,21–24,27,28] (Figure 1b). Indeed, although the
[3,20] (Box 1 and Figure 1a). Similar to the activated state, slow oscillations induced by anaesthesia might not show
the intracellular correlate of the slow (<1 Hz) oscillation in all the dynamic variability of those present during normal
individual cortical neurons is also characterized by periods sleep, they capture many important aspects of natural
of irregular tonic firing and a membrane potential that sleep oscillations and have, therefore, proven to be an
rapidly fluctuates around 60 mV [5,6] (Figure 1a). Unlike indispensable tool [3]. During anaesthetic treatment, brain
the activated state, however, such periods are regularly activation can be mimicked by stimulation of the ascending
interrupted by large, stereotypical hyperpolarizations that arousal system [29–32]. For example, under ketamine–
occur in tandem with the depth-positive component of the xylazine anaesthesia, brief electrical stimulation of the
EEG slow waves [5,6] (Box 1 and Figure 1a). These distinct pedunculopontine tegmentum (PPT) typically induces a
depolarized and hyperpolarized periods are usually prolonged period (20 s) of desynchronized EEG activity
referred to as ‘up’ and ‘down’ states, respectively. Thus, that is paralleled by a continuous depolarized state of
even in this cursory evaluation of intracellular recordings, the membrane potential (Figure 1b), similar to that
it is apparent that the ‘activated’ state of wakefulness and observed during natural waking [30–32] (Figure 1a).
‘up’ states of the slow (<1 Hz) sleep oscillation correspond Again, broad qualitative similarities between this stimu-
to broadly similar membrane-potential dynamics. lation-induced ‘activated’ state and ‘up’ states of the slow
Interestingly, several anaesthetic regimens, including oscillation are clearly evident, with the PPT stimulation
urethane and ketamine–xylazine, can also lead to a slow seeming to ‘lock’ the membrane potential into an extended
oscillation in the EEG of various species [10,21–28], with ‘up’ state.
www.sciencedirect.com
336 Review TRENDS in Neurosciences Vol.30 No.7

Figure 1. Cortical and thalamic activity in the cat during slow-wave oscillations and activated states. (a) Intracellular activity of cortical neurons during natural waking and
sleep states. During wakefulness (right panel), the depth EEG (top trace) is desynchronized and the membrane potential fluctuates around a depolarized level close to
60 mV. During SWS (left panel), the activity alternates between depolarized periods (‘up’ states; red bars), similar to wakefulness, and hyperpolarizations (‘down’ states),
synchronous with depth EEG slow waves. Modified and reproduced, with permission, from [5]. (b) Transformation of ‘up’ and ‘down’ state dynamics to activated states
under ketamine–xylazine anaesthesia. Brief electrical stimulation of the PPT (indicated schematically in the cat brain on the right) induces a transition from typical ‘up’ and
‘down’ state dynamics to a long-lasting EEG-activated state (see arrow above trace), in which the membrane potential seems to become ‘locked’ in the ‘up’ state. Modified
and reproduced, with permission, from [31]. (c) Simultaneous recording of depth EEG, intracellular activity from a TRN neuron and extracellular multi-unit TC neuron firing
in the ventral lateral (VL) nucleus (further enlarged bottom left) during both slow waves (left panel) and a period leading to transient brain activation (right panel); data were
obtained from the cat. The inset (bottom left) illustrates the intracellular nature of the slow oscillation in TC neurons. Modified and reproduced, with permission, from [32].
Note the close similarity in TRN neuron activity and VL firing between slow-oscillation ‘up’ states and continuous brain activation. Note also that the reversed polarity of the
EEG recording in (b), compared with (a) and (c), is owing to it being obtained using a surface electrode and explained in the Glossary.

The increased stability of intracellular recordings during neurons [21,32,33] (Figure 1c) and correspond closely to the
the anaesthetic-induced slow oscillation has proved vital for activity of thalamic neurons observed during natural sleep.
assessing the behaviour of neurons in key subcortical struc- Thus, the slow oscillation ‘up’ states in TC and TRN neurons
tures that are crucially involved in influencing the dynamics are characterized by seemingly irregular tonic firing and/or
of sleeping and waking states [21,32,33]. Specifically, intra- a membrane potential that fluctuates around 60 mV,
cellular recordings from both thalamocortical (TC) and whereas the ‘down’ state is manifest as a rhythmically
thalamic reticular nucleus (TRN) neurons have revealed a recurring, stereotypical hyperpolarizing event (Figure 1c).
pattern of ‘up’ and ‘down’ states during the slow oscillation Also, in close similarity to cortical neurons, the activity of
that are similar to, and synchronized with, those in cortical both types of thalamic neurons during brain activation
www.sciencedirect.com
Review TRENDS in Neurosciences Vol.30 No.7 337

becomes apparently ‘locked’ in a depolarized state (Box 1c), correlation excursions (Figure 2b). However, these
being characterized by continuous firing and/or fluctuations fluctuations in coherence are local, because the excursions
of the membrane potential that are superficially indistin- of correlations are greatly diminished between more dis-
guishable from those that occur during slow-oscillation ‘up’ tant pairs of electrodes [34–37] (Figure 2b). Extremely
states [21,32,33] (Figure 1c). The qualitative similarities similar results are obtained for ‘up’ states of the slow
between the cellular correlates of the ‘activated’ state and oscillation during natural sleep (Figure 2a), indicating that
‘up’ states of the slow (<1 Hz) sleep oscillation therefore the spatiotemporal dynamics of these ‘up’ states, as viewed
seem to be a ubiquitous feature of neuronal dynamics in through the EEG and LFPs, are essentially indistinguish-
corticothalamic networks. able from those of wakefulness.
Extracellularly recorded cortical neurons also display
Similarities in cortical network dynamics between similar dynamics during activated states and slow-oscil-
the ‘activated’ state and ‘up’ states of the slow lation ‘up’ states. It has been known for some time that
oscillation the mean firing rate of cortical neurons during wakefulness
The similarity between activated states and the and SWS is in the same range [19] (Figure 2c), a fact that is
slow-oscillation ‘up’ state in the cortex is not only apparent especially evident if the ‘up’ states of the slow oscillation are
at the rank of single cells, but can also be found at the level specifically analysed [5,34]. However, the dynamic sim-
of the EEG and local field potentials (LFPs) (see Glossary). ilarity concerns not only the mean rate, but also the
First, the typical desynchronized EEG pattern of arousal temporal patterns of discharge and the respective timing
is evident locally in the EEG during ‘up’ states, both in between different cells. Although such entities are more
the naturally sleeping animal (Figure 1a) and during difficult to characterize, a convenient, albeit subjective, way
anaesthesia (Figure 1b). A second, and stronger, indication to appreciate such information is by transforming the spike
of the similarities comes from studies using multiple bipo- patterns into audio signals by assigning a specific note to
lar electrodes in awake and naturally sleeping cats [34], each neuron. Allowing for the presence of ’concerted silences’
whereby it was shown that the desynchronized EEG pat- during SWS, owing to the slow-oscillation ‘down’ states, the
terns that are present during wakefulness are not only melodies produced by wakefulness and SWS are remarkably
irregular temporally, but, more generally, are also charac- similar (such audio material can be downloaded from http://
terized by low and fluctuating spatiotemporal coherence www.archive.org/details/NeuronalTones).
in LFPs (Figure 2a). The LFP signals at neighbouring Importantly, the similarity between cortical ‘up’ states
electrodes (at a distance of 1 mm) alternate between and activated states not only is evident in EEG and LFP
periods of high and periods of low coherence (HC dynamics and single-cell firing, but also extends to
and LC, respectively; Figure 2a), as quantified by the their relationship. Performing wave-triggered averages

Figure 2. Similar dynamic states during wakefulness and the ‘up’ state of SWS. (a) LFP activity in the cat parietal cortex during wakefulness (left) and SWS (right). Four
electrodes (traces 1–4) were placed at cortical depth along a line in the anterior–posterior axis of the suprasylvian gyrus at an interelectrode distance of 1 mm. The red boxes
indicate a period during which LFP activity had HC between two of the adjacent electrodes, whereas the blue boxes indicate a period of LC. During SWS, ‘up’ states were
identified by locally desynchronized EEG activity (indicated by the red bar). Modified and reproduced, with permission, from [34]. (b) Local correlation dynamics in LFP
activity. Correlations between adjacent electrodes calculated in small time windows (100 ms) show fluctuations between high and low values. The colour boxes indicate the
degree of correlation during the corresponding periods delineated in (a). Distant electrodes (1 and 4) did not show any significant correlation. Modified and reproduced,
with permission, from [34]. (c) Correlation between unit firing and LFP activity. The wave-triggered average shows that the negative LFP deflections (detected using a
thresholding procedure [34]; see example, inset) are correlated to increased firing in both wakefulness and ‘up’ states of SWS. The control indicates the absence of relation
after randomizing spike times. Modified and reproduced, with permission, from [34]. (d) Measurement of conductance in waking and ‘up’ states of SWS by matching
intracellular activities to stochastic models. The general pattern of mean conductance (g0) and conductance fluctuations (s) values was similar in waking and SWS states,
but the absolute values were different. Modified and reproduced, with permission, from [38].

www.sciencedirect.com
338 Review TRENDS in Neurosciences Vol.30 No.7

of spiking activity, by averaging the periods of neuronal intrinsic nature of the slow oscillation in thalamic neurons
firing around the negative peaks of the LFP (Figure 2c), offers a unique opportunity to examine the similarities
reveals that the depth-negative EEG component is corre- between slow-oscillation ‘up’ states and a continuously
lated to an increased probability of unit firing, both in the depolarized state (as occurs in natural wakefulness and
waking state and during slow-oscillation ‘up’ states brain activation; Figure 1c), because these two scenarios
(Figure 2c). can be invoked at will by simply varying the amount of
From the recent intracellular investigation of SWS [5], steady, injected current during an intracellular recording
the membrane-potential dynamics of cortical neurons were (Figure 3a,b). For example, the distribution of interspike
decomposed into excitatory and inhibitory conductance intervals (ISIs) in a condition in which thalamic neurons
components [38]. Such an analysis demonstrates that, in fire continuously in the absence of injected current
the majority of cortical cells, inhibition is stronger than (Figure 3) and a condition in which they are hyperpolarized
excitation, at the level of both mean conductance and to produce intermittent epochs of firing during slow-oscil-
conductance variation, as quantified by the standard devi- lation ‘up’ states (Figure 3) can be compared. Such a
ation (s) of the conductance (Figure 2d). This pattern is comparison reveals that the mode and essential form of
reported for both wakefulness and the ‘up’ states of SWS. the distribution of ISIs are extremely similar, although
However, there is a significant difference between the their coefficient of variation (CV) is always larger for ‘up’-
absolute values measured during the two states, with state-generated firing episodes than the ‘activated’ state.
SWS generally showing higher conductance (Figure 2d). Interestingly, the typical rate of firing of TRN neurons
Nevertheless, both states have a qualitatively similar ratio (40 Hz; Figure 3b) under both these conditions is always
of excitatory-to-inhibitory conductance, in which both the much higher than that of TC neurons (10 Hz; Figure 3a)
mean inhibitory conductance and its associated fluctu- [9,43,45]. This, in turn, suggests a dominance of inhibitory
ations are larger, on average, compared with excitatory over excitatory activity in the thalamus during both brain
contributions (Figure 2d). This resemblance also extends to activation and the slow-oscillation ‘up’ state, in a similar
the spike-triggered average conductance patterns in the manner to that revealed for cortical neurons (see above and
two states [38]. Figure 2d).
Finally, additional indirect evidence that cortical In most TC neurons, continuous firing during sustained
slow-oscillation ‘up’ states and wakefulness stem from depolarization consists solely of tonic firing of single action
similar network dynamics comes from modelling studies. potentials (Figure 3a) and, correspondingly, the slow-oscil-
Biophysical models with realistic cellular properties and lation ‘up’ state also shows straightforward firing of single
synaptic interactions have simulated cortical slow-oscil- action potentials [9,43] (Figure 2a). In a small number of
lation patterns of ‘up’ and ‘down’ states that are owing to TC neurons, however, firing during sustained depolariz-
recurrent interactions between networks of excitatory and ation comprises rhythmic (3–15 Hz) high-threshold (HT)
inhibitory neurons [39–41]. These models have established bursting [9,43] (Figure 3a) and, correspondingly, the slow-
that minimal parameter changes enable the transition oscillation ‘up’ state also produces epochs of HT bursting
from ‘up’ and ‘down’ states of the slow oscillation to a (Figure 3a). Because these HT burst-generating neurons
sustained ‘up’ state, with electrophysiological features might function as pacemakers for wake-related a (8–13 Hz)
consistent with experimental measurements. Thus, sev- rhythms [47–49], a distinct possibility is that they also
eral independent lines of experimental and modelling function to generate brief epochs of a-frequency activity
evidence indicate that cortical network dynamics during during the ‘up’ state of the slow oscillation in SWS [50,51].
slow-oscillation ‘up’ states are extremely similar to those
present during wakefulness or brain activation. Slow-oscillation ‘up’ states as micro-wake
‘fragments’
Similarities between a persistently depolarized An assortment of data from the corticothalamic system has
state and the ‘up’ state of the slow oscillation in been presented that converges to establish that, at both
thalamocortical and thalamic reticular nucleus single-cell and neuronal-network levels, the fully activated
neurons brain state and slow-oscillation ‘up’ states are dynamically
Both cortical and thalamic slow (<1 Hz) oscillations can be similar. An attractive interpretation of this is that indi-
reproduced using in vitro models. In cortical slices, the slow vidual corticothalamic ‘up’ states provide micro-wake-like
oscillation is reliant on a modified, artificial cerebrospinal contexts that facilitate specific types of neuronal inter-
fluid, containing a reduced Ca2+ concentration, and is action. In particular, they might provide brief epochs of
generated primarily by network-dependent mechanisms network dynamics that aid the transfer of memory traces
[24,42] (see Glossary). In thalamic slices, by contrast, the between short-term storage sites in the hippocampus and
slow oscillation in both TC and TRN neurons is generated long-term memory space in the neocortex. Within this
mainly by intrinsic mechanisms [9,43–46] (see Glossary). framework, the subtle differences that exist between the
These oscillations have extremely similar properties to dynamics of continuous waking and slow-oscillation ‘up’
those observed in thalamic cells during natural sleep states might function to alter the direction of information
[45,47] and become apparent if activation of modulatory flow between cortical areas while maintaining individual
cortical input is mimicked through either electrical stimu- neurons in a depolarized and active state that is conducive
lation of corticofugal afferents or the persistent pharma- to information processing. This modified information
cological activation of metabotropic glutamate receptors streaming might also be aided by the sleep-related changes
that are postsynaptic to these afferents [9,43–46]. The that occur in various neuromodulator levels, particularly
www.sciencedirect.com
Review TRENDS in Neurosciences Vol.30 No.7 339

Figure 3. Close quantitative similarities between the ‘up’ state of the slow oscillation and a persistently depolarized state in TC and TRN neurons. (a) Left column: intrinsic
slow (<1 Hz) oscillations, as recorded intracellularly from tonically hyperpolarized TC neurons in the cat lateral geniculate nucleus in vitro, exhibiting either simple tonic
firing (i, 60 pA) or HT bursting (ii, 50 pA) during the ‘up’ state (see expanded sections, lower left). The corresponding histogram of the distribution of ISIs (lower right) is
computed from several consecutive ‘up’ states. Right column: spontaneous activity of the neurons depicted in the left column; however, in the absence of hyperpolarizing
current, the activity consists of continuous tonic firing (i) and HT bursting (ii; see expanded sections, lower left). The corresponding histogram of the distribution of ISIs
(lower right) reveals an extremely similar pattern to that observed during slow-oscillation ‘up’ states. Adapted, with permission, from [9,43]. (b) Left column: intrinsic slow
oscillation, recorded intracellularly from a tonically hyperpolarized ( 50 pA) cat TRN neuron in vitro, exhibiting high-frequency tonic firing during the ‘up’ state (see
expanded section, lower left). The corresponding histogram of the distribution of ISIs (lower right) is computed from several consecutive ‘up’ states. Right column:
spontaneous activity of the same neuron in the absence of hyperpolarizing current consists of continuous tonic firing (see expanded sections, lower left). Again, the
corresponding histogram of the distribution of ISIs (lower right) reveals an extremely similar pattern to that observed during slow-oscillation ‘up’ states. Adapted, with
permission, from [45].

acetylcholine (Ach) [52]. These hypotheses are certainly Because the dynamics of slow-oscillation ‘up’ states are
consistent with the results of recent behavioural studies so similar to those of continuous wakefulness, why do they
showing that firing sequences from cell assemblies in both not lead to conscious experience during SWS? One possib-
the neocortex and the hippocampus are compressed and ility is that current methods and approaches overlook
reactivated concurrently with generation of the cortical several vital variables and are simply too coarse to detect
slow-oscillation ‘up’ state [12,13,53–56]. They also link the essential differences between what constitutes con-
well with human studies showing the following: first, local scious and unconscious dynamics. Alternatively, it could
increases in the average power density of slow oscillations be that the differences in dynamics between the activated
follow specific learning tasks and are related to a sub- state and ‘up’ states (e.g. discrepancies in neuronal con-
sequent improvement in task performance [14]; and sec- ductance and CVs, and compression of firing sequences)
ond, artificially boosting slow oscillations through the are sufficient to determine whether full awareness takes
transcranial application of oscillating fields enhances the place or not. In this sense, perhaps the most obvious and
retention of hippocampus-dependent declarative mem- intuitive candidate to explain the behavioural disparity
ories [16]. between wakefulness and sleep is the increased level of
www.sciencedirect.com
340 Review TRENDS in Neurosciences Vol.30 No.7

inhibition present during the ‘up’ state compared with the


waking state (Figure 2d). Another possibility is that con-
scious awareness might require an uninterrupted stream
of ‘activated’ neuronal dynamics lasting more than the few
seconds of an ‘up’ state, perhaps so that certain types of
long-range neuronal interactions during high-frequency
(20–80 Hz) activity, which have been suggested to break
down during sleep [57], can be fully established and main-
tained [58]. This is partly backed up by noting that during
sleep the most lucid experiences occur during the rapid eye
movement (REM) phase, during which the intermittent
hyperpolarizations (i.e. ‘down’ states) of the slow oscillation
are absent [5]. However, it should also be noted here that
not all studies support the hypothesis that long-range
interactions are diminished during sleep [15].

T-type Ca2+ channel-mediated bursts in


thalamocortical neurons as a trigger for ‘up’ states
and synaptic plasticity
If individual ‘up’ states contain segments of prior
wake-related dynamics, an attractive hypothesis is that
these segments are determined and selected during wake-
fulness through ongoing remodelling of cortical [59] or cor-
ticothalamic attractors by sensory input [18]. Such
attractors might then be preferentially activated in sleep Figure 4. Key events of the slow-oscillation cycle in thalamic and cortical neurons
during the slow-oscillation ‘up’ state [18], particularly in and their potential relation to plasticity. During the positive phase of the depth EEG
slow wave (top trace), neurons in both the thalamus (bottom two traces) and the
response to a strong thalamic signal, which is a highly cortex (second trace from the top) are silent. This scenario is terminated by
effective way to trigger internally defined cortical ensemble powerful T-type Ca2+ channel-dependent bursts in TC neurons. In turn, these bring
dynamics [60,61]. This is noteworthy because a conspicuous about widespread neuronal synchronization and Ca2+ entry, which provide
advantageous conditions for inducing cellular and synaptic plasticity. The degree
property of the slow oscillation in TC neurons is the presence to which cortical and thalamic neurons are susceptible to plasticity is indicated by
of a robust T-type Ca2+ channel-mediated burst at the the intensity of shading in the bar at the top of the figure. Although the opportunity
commencement of each ‘up’ state [9,21,32,33,43] (Box 1 for plasticity is probably most evident at the commencement of the ‘up’ state, it
might also persist later into the ‘up’ state, whereby wake-related neuronal activity
and Figure 4). Because these highly prominent TC neuron might be replayed and consolidated. Average cortical EEG and neuron activity
bursts invariably precede the ‘up’ state transitions in cor- modified and reproduced, with permission, from [34].
tical neurons in vivo [33], they might be the key network
triggers that are ultimately responsible for initiating a new
epoch of reactivated corticothalamic dynamics. Supporting activity during the early part of the ‘up’ state might further
evidence for this comes from the finding that in cortical contribute to bringing about cellular and synaptic
networks that lack thalamic input, i.e. cortical slabs [62] or plasticity. These suggestions are consistent with the
slices [24], the slow-oscillation ‘down’ state is considerably demonstrations that the slow-oscillation ‘up’ state is cru-
longer than that if thalamic input is present. In addition, cially related to memory consolidation in humans [15] and
recordings in the intact brain demonstrate that the firing cell-assembly reactivation in rats [13,53–56].
rate of a large proportion of cortical neurons is transiently
elevated at the start of each ‘up’ state [27,34] (Figure 4), Concluding remarks
which is consistent with a robust, stereotypical excitatory Electrophysiological and modelling data show that ‘up’
input. Thus, although spontaneous ‘up’ and ‘down’ states states of the slow (<1 Hz) sleep oscillation are dynamically
can emerge in isolated cortical networks [24,42] (Box 1), equivalent to the activated state of wakefulness. This is in
there is strong evidence that in the intact brain cortical ‘up’ agreement with several behavioural investigations, which
states might be triggered by thalamic input. indicate that waking activities are replayed, and possibly
In addition to functioning as network triggers, T-type consolidated, during SWS. The prominent T-type Ca2+
Ca2+ channel-mediated bursts might also have an import- channel-mediated bursts in TC neurons might function
ant role in facilitating plasticity. The Ca2+ entry associated as key network triggers that ensure the synchronous start
with these bursts, both directly (i.e. in thalamic neurons) of slow-oscillation ‘up’ states across related cortical terri-
and indirectly (i.e. in cortical neurons), could open the gate tories. The highly synchronized and increased firing that is
to subsequent modifications of synaptic strength and/or present in the initial portion of the ‘up’ state in different
intrinsic excitability (Figure 4), as previously proposed for neuronal components of corticothalamic networks might
sleep spindle oscillations (Box 1) in the cortex [63,64]. represent an important element in bringing about cellular
Moreover, the large increase of firing that occurs at the and synaptic plasticity.
start of the ‘up’ state in cortical neurons and its associated Clearly, several outstanding questions remain. First,
LFPs are synchronous over distances of several milli- why are the corticothalamic dynamics of slow-oscillation
metres [34]. This highly synchronous nature of neuronal ‘up’ states so similar to those of continuous wakefulness
www.sciencedirect.com
Review TRENDS in Neurosciences Vol.30 No.7 341

and why, considering this similarity, is the slow oscillation 22 Steriade, M. et al. (1993) Intracellular analysis of relations between the
slow (<1 Hz) neocortical oscillation and other sleep rhythms of the
associated with a lack of consciousness? Second, what are
electroencephalogram. J. Neurosci. 13, 3266–3283
the precise neuronal rudiments that define a lack of con- 23 Steriade, M. et al. (1993) A novel slow (<1 Hz) oscillation of neocortical
sciousness during sleep? Are our techniques too coarse to neurons in vivo: depolarizing and hyperpolarizing components.
determine this or is it simply that ‘up’ states are too short to J. Neurosci. 13, 3252–3265
sustain conscious experience? Third, what are the basic 24 Sanchez-Vives, M.V. and McCormick, D.A. (2000) Cellular and
network mechanisms of rhythmic recurrent activity in neocortex.
cellular mechanisms underlying SWS-specific learning
Nat. Neurosci. 3, 1027–1034
and memory consolidation? 25 Bokor, H. et al. (2005) Selective GABAergic control of higher-order
thalamic relays. Neuron 45, 929–940
Acknowledgements 26 Hahn, T.T. et al. (2006) Phase-locking of hippocampal interneurons’
This review is dedicated to the memory of Mircea Steriade. Work in our membrane potential to neocortical up–down states. Nat. Neurosci. 9,
laboratories is supported by The Wellcome Trust (grants 71436, 78311 1359–1361
and 78403 to V.C. and S.W.H. and the CNRS, Human Frontier Science 27 Haider, B. et al. (2006) Neocortical network activity in vivo is generated
Program and European Community (A.D. and M.R.). Additional through a dynamic balance of excitation and inhibition. J. Neurosci. 26,
information regarding other published work from our laboratories is 4535–4545
available at http://www.thalamus.org.uk and http://cns.iaf.cnrs-gif.fr. 28 Isomura, Y. et al. (2006) Integration and segregation of activity in
entorhinal–hippocampal subregions by neocortical slow oscillations.
References Neuron 52, 871–882
1 Hobson, J.A. (2005) Sleep is of the brain, by the brain and for the brain. 29 Steriade, M. et al. (1991) Network modulation of a slow intrinsic
Nature 437, 1254–1256 oscillation of cat thalamocortical neurons implicated in sleep delta
2 Achermann, P. and Borbely, A.A. (1997) Low-frequency (<1 Hz) waves: cortically induced synchronization and brainstem cholinergic
oscillations in the human sleep electroencephalogram. Neuroscience suppression. J. Neurosci. 11, 3200–3217
81, 213–222 30 Steriade, M. et al. (1993) Cholinergic and noradrenergic modulation
3 Steriade, M. and Amzica, F. (1998) Slow sleep oscillation, rhythmic K- of the slow (approximately 0.3 Hz) oscillation in neocortical cells.
complexes, and their paroxysmal developments. J. Sleep Res. 7 (Suppl J. Neurophysiol. 70, 1385–1400
1), 30–35 31 Rudolph, M. et al. (2005) Characterization of synaptic conductances
4 Molle, M. et al. (2002) Grouping of spindle activity during slow and integrative properties during electrically induced EEG-activated
oscillations in human non-rapid eye movement sleep. J. Neurosci. states in neocortical neurons in vivo. J. Neurophysiol. 94, 2805–2821
22, 10941–10947 32 Steriade, M. et al. (1996) Synchronization of fast (30–40 Hz) spon-
5 Steriade, M. et al. (2001) Natural waking and sleep states: a view from taneous cortical rhythms during brain activation. J. Neurosci. 16,
inside neocortical neurons. J. Neurophysiol. 85, 1969–1985 392–417
6 Timofeev, I. et al. (2001) Disfacilitation and active inhibition in the 33 Contreras, D. and Steriade, M. (1995) Cellular basis of EEG
neocortex during the natural sleep–wake cycle: an intracellular study. slow rhythms: a study of dynamic corticothalamic relationships.
Proc. Natl. Acad. Sci. U. S. A. 98, 924–929 J. Neurosci. 15, 604–662
7 Lee, J. et al. (2004) Lack of delta waves and sleep disturbances during 34 Destexhe, A. et al. (1999) Spatiotemporal analysis of local field
non-rapid eye movement sleep in mice lacking a1G-subunit of T-type potentials and unit discharges in cat cerebral cortex during natural
calcium channels. Proc. Natl. Acad. Sci. U. S. A. 101, 18195–18199 wake and sleep states. J. Neurosci. 19, 4595–4608
8 Sirota, A. et al. (2003) Communication between neocortex and 35 Eckhorn, R. et al. (1988) Coherent oscillations: a mechanism of feature
hippocampus during sleep in rodents. Proc. Natl. Acad. Sci. U. S. A. linking in the visual cortex? Multiple electrode and correlation
100, 2065–2069 analyses in the cat. Biol. Cybern. 60, 121–130
9 Zhu, L. et al. (2006) Nucleus- and species-specific properties of the slow 36 Gray, C.M. and Singer, W. (1989) Stimulus-specific neuronal
(<1 Hz) sleep oscillation in thalamocortical neurons. Neuroscience 141, oscillations in orientation columns of cat visual cortex. Proc. Natl.
621–636 Acad. Sci. U. S. A. 86, 1698–1702
10 Wolansky, T. et al. (2006) Hippocampal slow oscillation: a novel EEG 37 Steriade, M. and Amzica, F. (1996) Intracortical and corticothalamic
state and its coordination with ongoing neocortical activity. J. coherency of fast spontaneous oscillations. Proc. Natl. Acad. Sci. U. S. A.
Neurosci. 26, 6213–6229 93, 2533–2538
11 Eschenko, O. et al. (2006) Elevated sleep spindle density after learning 38 Rudolph, M. et al. Inhibition determines membrane potential dynamics
or after retrieval in rats. J. Neurosci. 26, 12914–12920 and controls action potential generation in awake and sleeping cat
12 Ji, D. and Wilson, M.A. (2007) Coordinated memory replay in the visual cortex. J. Neurosci. (in press) (http://www.jneurosci.org/)
cortex and hippocampus during sleep. Nat. Neurosci. 10, 100–107 39 Bazhenov, M. et al. (2002) Model of thalamocortical slow-wave sleep
13 Lee, A.K. and Wilson, M.A. (2002) Memory of sequential experience in oscillations and transitions to activated States. J. Neurosci. 22, 8691–
the hippocampus during slow wave sleep. Neuron 36, 1183–1194 8704
14 Huber, R. et al. (2004) Local sleep and learning. Nature 430, 78–81 40 Compte, A. et al. (2003) Cellular and network mechanisms of slow
15 Molle, M. et al. (2004) Learning increases human electroencephalo- oscillatory activity (<1 Hz) and wave propagations in a cortical
graphic coherence during subsequent slow sleep oscillations. Proc. network model. J. Neurophysiol. 89, 2707–2725
Natl. Acad. Sci. U. S. A. 101, 13963–13968 41 Alvarez, F.P. and Destexhe, A. (2004) Simulating cortical network
16 Marshall, L. et al. (2006) Boosting slow oscillations during sleep activity states constrained by intracellular recordings. Neurocom-
potentiates memory. Nature 444, 610–613 puting 58, 285–290
17 Huber, R. et al. (2006) Arm immobilization causes cortical plastic 42 Cunningham, M.O. et al. (2006) Neuronal metabolism governs cortical
changes and locally decreases sleep slow wave activity. Nat. network response state. Proc. Natl. Acad. Sci. U. S. A. 103, 5597–5601
Neurosci. 9, 1169–1176 43 Hughes, S.W. et al. (2002) Cellular mechanisms of the slow (<1 Hz)
18 Battaglia, F.P. et al. (2005) Firing rate modulation: a simple statistical oscillation in thalamocortical neurons in vitro. Neuron 33, 947–958
view of memory trace reactivation. Neural Netw. 18, 1280–1291 44 Crunelli, V. et al. (2005) The ‘window’ T-type calcium current in brain
19 Hobson, J.A. and McCarley, R.W. (1971) Cortical unit activity in sleep dynamics of different behavioural states. J. Physiol. 562, 121–129
and waking. Electroencephalogr. Clin. Neurophysiol. 30, 97–112 45 Blethyn, K.L. et al. (2006) Neuronal basis of the slow (<1 Hz) oscillation
20 Amzica, F. and Steriade, M. (1998) Cellular substrates and laminar in neurons of the nucleus reticularis thalami in vitro. J. Neurosci. 26,
profile of sleep K-complex. Neuroscience 82, 671–686 2474–2486
21 Steriade, M. et al. (1993) The slow (<1 Hz) oscillation in reticular 46 Crunelli, V. et al. (2006) Thalamic T-type Ca2+ channels and NREM
thalamic and thalamocortical neurons: scenario of sleep rhythm sleep. Cell Calcium 40, 175–190
generation in interacting thalamic and neocortical networks. 47 Hughes, S.W. et al. (2004) Synchronized oscillations at a and W
J. Neurosci. 13, 3284–3299 frequencies in the lateral geniculate nucleus. Neuron 42, 253–268

www.sciencedirect.com
342 Review TRENDS in Neurosciences Vol.30 No.7

48 Hughes, S.W. and Crunelli, V. (2005) Thalamic mechanisms of EEG a 57 Massimini, M. et al. (2005) Breakdown of cortical effective connectivity
rhythms and their pathological implications. Neuroscientist 11, during sleep. Science 309, 2228–2232
357–372 58 Roelfsema, P.R. et al. (1997) Visuomotor integration is associated
49 Hughes, S.W. and Crunelli, V. (2006) Just a phase they’re going with zero time-lag synchronization among cortical areas. Nature 385,
through: the complex interaction of intrinsic high-threshold bursting 157–161
and gap junctions in the generation of thalamic a and W rhythms. Int. J. 59 Cossart, R. et al. (2003) Attractor dynamics of network UP states in the
Psychophysiol. DOI: 10.1016/j.ijpsycho.2006.08.004 neocortex. Nature 423, 283–288
50 MacFarlane, J.G. et al. (1996) Periodic K-a sleep EEG activity and 60 MacLean, J.N. et al. (2005) Internal dynamics determine the cortical
periodic limb movements during sleep: comparisons of clinical features response to thalamic stimulation. Neuron 48, 811–823
and sleep parameters. Sleep 19, 200–204 61 Amzica, F. and Steriade, M. (2001) Cellular substrates and laminar
51 Karadeniz, D. et al. (2000) EEG arousals and awakenings in relation profile of sleep K-complex. Neuroscience 82, 671–686
with periodic leg movements during sleep. J. Sleep Res. 9, 273–277 62 Timofeev, I. et al. (2000) Origin of slow cortical oscillations in
52 Hasselmo, M.E. (1999) Neuromodulation: acetylcholine and memory deafferented cortical slabs. Cereb. Cortex 10, 1185–1199
consolidation. Trends Cogn. Sci. 3, 351–359 63 Contreras, D. et al. (1997) Intracellular and computational
53 Siapas, A.G. and Wilson, M.A. (1998) Coordinated interactions characterization of the intracortical inhibitory control of
between hippocampal ripples and cortical spindles during slow-wave synchronized thalamic inputs in vivo. J. Neurophysiol. 78, 335–350
sleep. Neuron 21, 1123–1128 64 Sejnowski, T.J. and Destexhe, A. (2000) Why do we sleep? Brain Res.
54 Nadasdy, Z. et al. (1999) Replay and time compression of recurring 886, 208–223
spike sequences in the hippocampus. J. Neurosci. 19, 9497–9507 65 Contreras, D. et al. (1996) Control of spatiotemporal coherence of a
55 Battaglia, F.P. et al. (2004) Hippocampal sharp wave bursts coincide thalamic oscillation by corticothalamic feedback. Science 274, 771–774
with neocortical ‘‘up-state’’ transitions. Learn. Mem. 11, 697–704 66 Contreras, D. and Steriade, M. (1996) Spindle oscillation in cats: the
56 Molle, M. et al. (2005) Hippocampal sharp wave-ripples linked to slow role of corticothalamic feedback in a thalamically generated rhythm.
oscillations in rat slow-wave sleep. J. Neurophysiol. 96, 62–70 J. Physiol. 490, 159–179

Five things you might not know about Elsevier


1.
Elsevier is a founder member of the WHO’s HINARI and AGORA initiatives, which enable the
world’s poorest countries to gain free access to scientific literature. More than 1000 journals,
including the Trends and Current Opinion collections and Drug Discovery Today, are now available
free of charge or at significantly reduced prices.

2.
The online archive of Elsevier’s premier Cell Press journal collection became freely available in
January 2005. Free access to the recent archive, including Cell, Neuron, Immunity and Current
Biology, is available on ScienceDirect and the Cell Press journal sites 12 months after articles are
first published.

3.
Have you contributed to an Elsevier journal, book or series? Did you know that all our authors are
entitled to a 30% discount on books and stand-alone CDs when ordered directly from us? For
more information, call our sales offices:

+1 800 782 4927 (USA) or +1 800 460 3110 (Canada, South and Central America)
or +44 (0)1865 474 010 (all other countries)

4.
Elsevier has a long tradition of liberal copyright policies and for many years has permitted both the
posting of preprints on public servers and the posting of final articles on internal servers. Now,
Elsevier has extended its author posting policy to allow authors to post the final text version of
their articles free of charge on their personal websites and institutional repositories or websites.

5.
The Elsevier Foundation is a knowledge-centered foundation that makes grants and contributions
throughout the world. A reflection of our culturally rich global organization, the Foundation has,
for example, funded the setting up of a video library to educate for children in Philadelphia,
provided storybooks to children in Cape Town, sponsored the creation of the Stanley L. Robbins
Visiting Professorship at Brigham and Women’s Hospital, and given funding to the 3rd
International Conference on Children’s Health and the Environment.

www.sciencedirect.com

You might also like