Download Spontaneous activity in developing sensory circuits

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

History of neuroimaging wikipedia , lookup

Binding problem wikipedia , lookup

Emotional lateralization wikipedia , lookup

Central pattern generator wikipedia , lookup

Theta wave wikipedia , lookup

Executive functions wikipedia , lookup

Electroencephalography wikipedia , lookup

Sensory substitution wikipedia , lookup

Neuromarketing wikipedia , lookup

Clinical neurochemistry wikipedia , lookup

Holonomic brain theory wikipedia , lookup

Brain–computer interface wikipedia , lookup

Microneurography wikipedia , lookup

Environmental enrichment wikipedia , lookup

Activity-dependent plasticity wikipedia , lookup

Synaptic gating wikipedia , lookup

Time perception wikipedia , lookup

Aging brain wikipedia , lookup

Affective neuroscience wikipedia , lookup

Connectome wikipedia , lookup

Cortical cooling wikipedia , lookup

Optogenetics wikipedia , lookup

Human brain wikipedia , lookup

Functional magnetic resonance imaging wikipedia , lookup

Development of the nervous system wikipedia , lookup

Neuroesthetics wikipedia , lookup

Evoked potential wikipedia , lookup

Premovement neuronal activity wikipedia , lookup

Biology of depression wikipedia , lookup

Magnetoencephalography wikipedia , lookup

Cognitive neuroscience of music wikipedia , lookup

Eyeblink conditioning wikipedia , lookup

Neuroeconomics wikipedia , lookup

Feature detection (nervous system) wikipedia , lookup

Neuroplasticity wikipedia , lookup

Neuropsychopharmacology wikipedia , lookup

Neural oscillation wikipedia , lookup

Cerebral cortex wikipedia , lookup

Metastability in the brain wikipedia , lookup

Neural correlates of consciousness wikipedia , lookup

Spike-and-wave wikipedia , lookup

Transcript
YNIMG-09254; No. of pages: 10; 4C:
NeuroImage xxx (2012) xxx–xxx
Contents lists available at SciVerse ScienceDirect
NeuroImage
journal homepage: www.elsevier.com/locate/ynimg
1
Review
Q1 4
Matthew Colonnese a, b, c,⁎, Rustem Khazipov a, b, d
O
a
INSERM, U29, INMED, Marseille 13009, France
Aix Marseille Université, Faculté des Sciences, Marseille, F-13000, France
Department of Pharmacology and Physiology, Institute for Neuroscience, George Washington University, Washington DC 20037, USA
d
Kazan Federal University, Kazan, Russia
b
c
R
O
5
6
Q2 7
8
F
Q3 3
Spontaneous activity in developing sensory circuits: Implications for resting
state fMRI
2
9
i n f o
a b s t r a c t
P
a r t i c l e
Article history:
Accepted 13 February 2012
Available online xxxx
D
The immature brain spontaneously expresses unique patterns of electrical activity that are believed to contribute to the development of neuronal networks. Certain electrographic features of this activity, particularly
modulation on an infraslow time scale, resemble activity patterns observed in the mature brain at ‘rest’,
loosely defined as the absence of an investigator imposed task. However, it is not clear whether the immature
activity patterns observed at rest are precursors of the spontaneous neuronal activity that forms resting state
networks in the adult. Here, we review recent studies that have explored the generative mechanisms of resting state activity during development in the primary sensory systems of premature human neonates and
neonatal rodents. The remarkable hypothesis suggested by this work is that while resting state activity during the pre- and possibly near-term period can bear superficial resemblance to adult activity it is fundamentally different in terms of function and origin. During early development spontaneous thalamocortical activity
in primary sensory regions is determined largely by transitory generators in the sensory periphery. This is in
contrast to the adult, where spontaneous activity generated within thalamocortex, particularly by corticocortical connections, dominates. We therefore suggest a conservative interpretation of developmental mapping studies which are based on indirect measurement of activity (e.g. fMRI), or on the partitioning of EEG
frequency using bands derived from adult studies. The generative mechanisms for brain activity at early
ages are likely different from those of adults, and may play very different roles; for example in circuit formation as opposed to attention.
© 2012 Published by Elsevier Inc.
T
E
Keywords:
Development
Resting state
Preterm
Neonatal
Spindle burst
Spontaneous activity
R
E
C
10
11
12
13
15
14
16
17
18
19
20
21
22
23
N
C
O
48
49
50
51
52
53
54
55
56
57
58
Q1459
60
Contents
Resting state activity in neonates and adults . . . . . . . . . . .
Discontinuous temporal organization of early activity . . . . . . .
Unique activity patterns in preterm infants . . . . . . . . . . . .
Appearance . . . . . . . . . . . . . . . . . . . . . . . .
Modulation . . . . . . . . . . . . . . . . . . . . . . . .
Developmental separation . . . . . . . . . . . . . . . . .
Generative mechanisms . . . . . . . . . . . . . . . . . .
Peripheral dominance of resting activity during early development
Development of intrinsic thalamocortical activities . . . . . . . .
Early cortical function and implications for resting state fMRI . . .
Summary . . . . . . . . . . . . . . . . . . . . . . . . . . .
Uncited references . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . .
U
44
47
46
43
42
R
45
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
61
⁎ Corresponding author at: INSERM, U29, INMED, Marseille 13009, France.
E-mail address: [email protected] (M. Colonnese).
1053-8119/$ – see front matter © 2012 Published by Elsevier Inc.
doi:10.1016/j.neuroimage.2012.02.046
Please cite this article as: Colonnese, M., Khazipov, R., Spontaneous activity in developing sensory circuits: Implications for resting state fMRI,
NeuroImage (2012), doi:10.1016/j.neuroimage.2012.02.046
0
0
0
0
0
0
0
0
0
0
0
0
0
M. Colonnese, R. Khazipov / NeuroImage xxx (2012) xxx–xxx
80
81
Q4 82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
Q5 107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
C
F
O
78
79
E
76
77
R
74
75
R
72
73
O
70
71
C
68
69
N
67
U
65
66
R
O
The cerebral cortex is never quiet (Fox and Raichle, 2007). Even primary sensory regions are consistently active in the absence of sensory
input, including during states of unconsciousness such as non-REM
sleep and anesthesia (Bianciardi et al., 2009; Hasenstaub et al., 2007;
Leopold and Logothetis, 2003). Neurophysiologists have a tradition of
labeling such non-task related activity ‘spontaneous’ or even ‘background’, while similar fluctuations have been categorized as ‘resting
state’ by the human fMRI community. The origins and functions of
this activity are a mystery, but are thought to include replay of sensory
information in cortical circuits during sleep and attention, visual
learning and ongoing cogitation during waking (Berkes et al., 2011;
Haider and McCormick, 2009; Ji and Wilson, 2007). And, resting state
activity fluctuations have themselves become a powerful tool in
defining the functional connections between cortical and sub-cortical
regions. Measuring the co-modulation of activity in multiple regions
with fMRI or EEG during ‘rest’, i.e. in the absence of directed goals, has
revealed multiple independent networks linked to sensory processing,
movement, and higher functions such as attention. Perhaps most striking
is the discovery of ‘default-mode’ networks inversely correlated with
directed behavior and perceptual tasks (Raichle and Snyder, 2007).
Understanding when these networks form is an important goal, because their disruption during development is predicted to be causally
linked to the onset of neurological disorders. The question has two components: when do the anatomical connections that link brain regions
together form, and when are the network properties sufficiently mature
to allow for mature function? This latter question is critically important
because network properties change constantly during development.
Animal models have shown that during the initial formation of neural
circuits a number of transitory properties exist to generate spontaneous
activity in the local circuits, despite a poverty of local interconnectivity
and adult functional connectivity (Blankenship and Feller, 2010). For
example, isolated slices of rodent cortex and hippocampus exhibit multiple stages of network activity during the first postnatal weeks (Allene
et al., 2008; Moody and Bosma, 2005). In addition, both the retina and
cochlea generate transitory activity bursts that synchronize firing in
local neuronal populations on an infraslow timescale (Demas et al.,
2003; Tritsch et al., 2007; Wong, 1999). This process has been studied
extensively in the visual system, where genetic and pharmacological
manipulation of spontaneous retinal activity has been shown to be critical for visual map formation in central visual structures including the
cortex (Huberman et al., 2008). Sensorimotor systems also engage in
a similar process, but instead of generating activity spontaneously in
the sense organ, spontaneous bursts of activity in the spinal cord
cause myclonic jerks in the various muscles, resulting in ‘spontaneous’
sensory input that is carried to the cortex (Khazipov et al., 2004;
Kreider and Blumberg, 2000).
In short, data from animal models predict significant ‘resting-state’
activity during early developmental time periods equivalent to the
fetal stage in humans. This review examines the evidence that this activity is not equivalent to resting-state modulations in adults, either in
terms of generative mechanisms or function. This is important because
understanding the development of resting state networks requires not
just a demonstration of co-modulation of various brain regions, but
also an understanding of the generative mechanisms of these modulations, which are likely to change dramatically during development.
Humans are born in the midst of a surge of cortical synaptogenesis
that will continue throughout the first year after birth (Huttenlocher
and Dabholkar, 1997). This synaptic burst is coincident with a rapid increase in the continuity of cortical activity and a maturation of activity
patterns, as many EEG grapho-elements disappear in the first couple of
months, and adult patterns of activity including sleep spindles, delta
waves and strong gamma-band modulations emerge within the first
year (Andre et al., 2010; Curzi-Dascalova, 1977; Ellingson and Peters,
1980). As might be expected, these rapid changes in synaptic density,
P
63
64
connectivity and electrical activity are accompanied by rearrangements
in resting state cortical networks as assayed by fMRI. Resting state networks, organized locally around hubs in sensory and motor cortex,
have been demonstrated in term infants (Fransson et al., 2011), as well
as preterm infants (Doria et al., 2010; Smyser et al., 2010). By contrast,
resting state networks involving distant regions, particularly frontal
and associative regions such as the default state and attention networks,
develop later than local networks. One study identified the adult complement of resting state networks by term (Doria et al., 2010), while other
studies observed an incomplete default state network at term
(Fransson et al., 2007, 2011; Gao et al., 2009; Smyser et al., 2010).
These data suggest that some important property of sensory networks forms in utero, while the development of associational networks
is likely part of the rapid synaptogenesis and maturation of cortical
activity patterns observed postnatally. Thus a close examination of activity in primary sensory cortex during the fetal period will be necessary
to determine to what extent the resting state fMRI detects adult-like
functional networks or immature activity patterns, and what this
would mean for the processing capacity of the infant brain.
Determining the patterns of activity expressed in the fetal brain is
a challenging task. Due to the technical difficulty of performing EEG
and MEG through the uterine wall, the main approach consists of
scalp EEG recordings from premature neonates. One should question
whether cortical activity in premature neonates is the same as in the
fetus. Certainly, the extra-uterine environment is very different, and
the function of many systems (cardiovascular, respiratory, and digestive) undergo significant development after birth regardless of gestational age. However, the major characteristics of preterm EEG, such as
those used for clinical monitoring, are primarily determined by gestational, and not postnatal, age (Andre et al., 2010); the length of time
ex utero does not exert a large effect on the gross patterns of brain
activity maturation. Moreover, the temporal organization of spontaneous activity as shown by fetal MEG studies is remarkably consistent
with data obtained from gestational age matched premature neonates
(Eswaran et al., 2004, 2007). A number of caveats apply, however. The
development of a number of neurological markers is accelerated by preterm birth, while others are delayed or unaffected (Ellingson and Peters,
1980; Smyser et al., 2010) and these can vary depending on the extrauterine experience (Guzzetta et al., 2009). Furthermore, the fetus,
under the influence of maternal hormones and lack of stimuli, is maintained in a sleep-promoting environment. Therefore, scalp EEG from
healthy premature neonates should be considered as an approximation
of fetal brain activity, and hopefully further technical developments will
provide further verification for this model.
A mechanistic understanding of resting state network development
requires appropriate animal models. Pioneering studies describing the
phenomenology of EEG development were performed in fetal and neonatal pigs, sheep, guinea pigs, cats, rabbits, rats and dogs (Bergstrom et
al., 1962; Bernhard et al., 1959; Charles and Fuller, 1956; Crain, 1952;
Deza and Eidelberg, 1967; Fox, 1967; Jouvet-Mounier et al., 1970;
Marty and Thomas, 1963; Pampiglione, 1977; Petersen et al., 1964;
Rose and Ellingson, 1968; Scherrer and Oeconomos, 1954; Tuge et al.,
1960; Yoshii and Tsukiyama, 1951). These studies led to conclusion
that EEG pattern maturation is similar between species and determined
more by relative maturation than by time of birth. In all species, at first
appearance EEG activity consists of bursts of activity lasting around 1 s
with intervening periods of electrical silence that can last up to 10 s.
Interhemispheric synchrony is lacking. No changes associated with
the sleep-wake cycle can be detected. As described by Ellingson
(1970) EEG patterns then undergo a series of universal developmental
trends in the postnatal EEG: (1) low frequency activity becomes less
abundant with increasing age; (2) frequency of rhythmic (alpha- or
alpha-like) activity during wakefulness increases with age; (3) interhemispheric synchrony improves with age; (4) overall amplitude increases during infancy and then declines somewhat to maturity; and
(5) complexity of sleep patterns increases with age (Ellingson, 1970).
D
Resting state activity in neonates and adults
T
62
E
2
Please cite this article as: Colonnese, M., Khazipov, R., Spontaneous activity in developing sensory circuits: Implications for resting state fMRI,
NeuroImage (2012), doi:10.1016/j.neuroimage.2012.02.046
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
M. Colonnese, R. Khazipov / NeuroImage xxx (2012) xxx–xxx
214
215
216
217
218
219
220
Q6 221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
246
247
248
249
250
251
252
253
254
255
256
C
212
213
E
316
We have considered the differences in the nature of the infraslow
oscillations observed at rest during early development and in adults,
namely the alternation between network silence and activity that characterizes the immature brain, as opposed to waxing and waning of
317
F
Unique activity patterns in preterm infants
O
210
211
R
208
209
R
206
207
N
C
O
204
205
U
202
203
R
O
The adult brain is constantly active, though the patterns and amount
of this activity are modulated differentially between brain networks by
a number of factors including vigilance state, mental task and sensory
input (Steriade and McCarley, 2005). The basis of functional connectivity mapping appears to be the waxing and waning of ongoing activity on
an ‘infraslow’ timescale (0.01–0.2 Hz), rather than the cessation and reinduction of activity itself (Baria et al., 2011; Jann et al., 2010; Laufs et
al., 2003; Sadaghiani et al., 2010; Vincent et al., 2007; Zuo et al.,
2010). Since the first recordings of human neonates using scalp electrographic recordings by Dreyfus-Brisac and Monod (1965) and DreyfusBrisac et al., (1956), a pronounced infraslow modulation of activity in
preterm infants has also been noted. However, close examination of
this oscillation reveals that it is of a fundamentally different character
200
201
P
244
245
199
257
258
D
Discontinuous temporal organization of early activity
197
198
from adult: it arises not from the modulation of ongoing activity or
brain states as in adults, but from alternating periods of complete network silence and activity. This temporal organization in preterms was
named tracé discontinu, and is most prominent during the second half
of gestation, particularly during quiet sleep. During tracé discontinu,
the cortical EEG is organized in intermittent bursts separated by periods
of isoelectric EEG that could last for tens of seconds (Dreyfus-Brisac and
Larroche, 1971). With maturation, the flat periods between bursts become progressively shorter, first during active (pre-REM) sleep and
then during quiet sleep. By 30 weeks post-conception periods of tracé
discontinu become interleaved with periods of low-voltage activity (indicative of continuous activity) observed between the bursts. These periods, so called tracé alternant, become more common, resulting in
longer periods of continuous activity. Continuous cortical activity is observed from 1 to 2 months after birth (Andre et al., 2010).
Thus infraslow modulation in preterm infants results from the apparent cessation of activity, while in adults it results from more modest
increases and decreases in ongoing activity. Recent rodent studies have
examined whether these isoelectric ‘silent’ periods result from true network silence, or from asynchronous neuronal activity that poorly registers at the scalp. Extracellular recordings of multiple unit activity in
somatosensory and visual cortex of rats showed that neurons fire virtually no action potentials during these isoelectric EEG epochs (Fig. 1). In
agreement with this result, patch-clamp recordings revealed that neurons stay relaxed at their resting membrane potential and receive
very few synaptic inputs during the isoelectric EEG periods
(Colonnese and Khazipov, 2010; Khazipov et al., 2004; Rochefort et al.,
2009). This means that isoelectric EEG epochs reflect a synchronous
neuronal silence which is never observed in a healthy adult. The closest
the adult cortex comes to such collective silence is during so called
‘down states’, observed during unconsciousness brought on by slowwave sleep or anesthesia. In adults, natural down states observed
during sleep last only on the order of hundreds of milliseconds
(Steriade and McCarley, 2005), whereas they can last for tens of seconds
in neonatal rats and preterm humans.
Oscillations between periods of neuronal depolarization associated
with synaptic barrages (called UP-states) and the DOWN state generates slow waves on EEG (Steriade et al., 1993), giving name to the
state of sleep when they are most commonly observed. In adults this
so-called ‘slow’ oscillation between up and down states is too rapid to
account for resting state connectivity (though the number or power of
such transitions over a 10 s period could in theory (Jann et al., 2010)).
In contrast, in preterm infants and neonatal rats a singular burst of activity against a background of network silence could conceivably contribute to infraslow modulations and resting state connectivity.
Why do immature networks produce such long-lasting periods of
network silence? Intracellular studies in adults indicate that down
states result when the level of mutual excitation in the cortical network
falls below the critical level necessary to maintain the UP-state, allowing neurons to return to their resting membrane potential (Contreras
et al., 1996; Sanchez-Vives and McCormick, 2000; Shu et al., 2003). Immature networks, which are much less densely connected, likely require much greater network activation before generating an up-state,
and thus spend much longer periods of time in network silence. Combined with the evidence outlined below that the majority of activity
that does occur in immature sensory cortex is related to thalamic
input, it appears that the default state of cortical networks during the
fetal period is network silence, rather than the continuous ‘spontaneous’ activity characteristic of healthy mature cortex.
E
243
195
196
T
241
242
Pampiglione (1977) also pointed out that “from the evolution of E.E.G.
features it seems that the underlying mechanisms during maturation
do not evolve according to a uniformly continuous process but proceed,
like the main E.E.G. features, in irregular steps, with plateaus lasting
days or weeks, followed by a more rapid change over a period of a
very few days before reaching the next plateau which then may last
several weeks or months before changing again into a final adult
phase. The evolution process therefore seems to be much more like a series of changes of uneven duration rather than a smooth curve.”
To move beyond these phenomenological observations and construct
a mechanistic understanding of the early activity patterns and their
physiological roles in brain development, recent efforts have focused
on neonatal rodents, and to a lesser extent ferrets, as a model system
for human EEG development. In rodents invasive recordings of brain activity, including intra-cellular recordings, can be combined with pharmacological and genetic manipulations to identify network mechanisms of
early activity generation (Khazipov and Luhmann, 2006). Rats and
mice are altricial; that is, they are born in an immature state. Although
it is difficult to provide exact comparisons between human and rodents,
the degree of rat cortical development on the day of birth (P0) can be
roughly compared to the stage of human cortex at mid-gestation
(Clancy et al., 2001). Multi-factor estimates based primarily on neurogenetic times (Clancy et al., 2007), and direct comparisons of neural activity
patterns in humans and rats (Colonnese et al., 2010) disagree on the rat
equivalent of term in humans. This is in part because there is likely no
single age equivalent useful for all brain regions, as even primates exhibit
profound heterogeneity between cortical and sub-cortical developmental times (Clancy et al., 2000). For hippocampus, term in human was suggested to correspond to postnatal day P5–7 in the rat (Avishai-Eliner et
al., 2002). For cortex however, both the timing of the surge in synaptogenesis in primary sensory regions, and changes in neural activity appear
to develop on a time course that diverges from other developmental
markers. Such heterochrony may have implications for resting state network development, as networks may become ‘functional’ at different developmental points in different species. For visual cortex, in terms of
function, we suggest that term in humans roughly corresponds to postnatal day P12-14 (roughly the time of eye opening) in the rat or mouse.
The invasive animal studies completed so far suggest that in primary
sensory systems neuronal activity during the preterm period is categorically different from adult activity, even when there are superficial similarities in terms of temporal modulation and frequencies. In particular,
while adult resting activity appears generated by the dense corticocortical interconnectivity modulated by the neuromodulatory systems
(Destexhe et al., 2003; Haider and McCormick, 2009), early sensory networks are driven largely by spontaneous activity in the periphery, interacting with thalamocortical network properties unique to the time
period. We will review this evidence below, focusing on two features:
the lack of continuous cortical–cortical activity, which causes a discontinuous pattern of resting activity, and the generation of unique activity
patterns in cortex during early development.
193
194
3
Please cite this article as: Colonnese, M., Khazipov, R., Spontaneous activity in developing sensory circuits: Implications for resting state fMRI,
NeuroImage (2012), doi:10.1016/j.neuroimage.2012.02.046
259
260 Q7
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283 Q8
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
318
319
320
4
M. Colonnese, R. Khazipov / NeuroImage xxx (2012) xxx–xxx
P6
∗
∗ ∗
∗
Quiet Inter-SAT
∗
unique rhythmic, oscillations and intermittent, fast events that are
expressed during restricted ages. These immature patterns are notable
for their poor modulation by specific behaviors or vigilance states. Because they are clearly different in structure and occurrence from adult
patterns, these events have been given unique names, which include
temporal theta bursts, encoches frontales, and delta brush (Andre et al.,
2010; Mizrahi et al., 2003). The earliest activity, detected at 24–25
post-menstrual weeks (the earliest age that reliable recordings are
made), appears in clinical EEG as intermittent ‘delta’ waves, either
smooth or with more rapid rhythms superimposed. By the seventh
month, these slow oscillations are reliably intermixed with rapid
rhythms, and are abundant in central, temporal and occipital regions,
fading first in central regions and lastly on occipital electrodes around
term.
Our understanding of the organization and development of early activity patterns is still evolving. Recent observations using direct-current
(DC) recordings to reveal activity modulation on long timescales show
that delta brushes and other rapid oscillations often occur sequentially,
contained within a negative potential of unusually large amplitude (up
to 800 μV; (Vanhatalo et al., 2002)). During standard EEG such ‘DC
shifts’ are not observed due to the use of conventional high-pass
(>0.5–1 Hz) filters. These studies suggest that the subdivision of various oscillations into discrete events such as individual delta brushes
should be questioned, and that they would be better considered a singular class of unique developmental events termed ‘spontaneous activity transients’ (SATs), which consist of slow and infraslow waves that
group rapid oscillations of various frequencies (Vanhatalo and Kaila,
2006). Evidence from animal studies, reviewed below, supports this
classification.
Wide Band EEG
500μV
40 s
200 μV
2s
P10
Bursting Inter-SAT
∗
∗
∗
500μV
O
F
40 s
200 μV
R
O
1s
Constant retinal
Activity
No retinal activity
∗
Continuous Inter-SAT
∗ ∗
∗
D
P13
P
500μV
40 s
E
500μV
Modulation
T
40 s
321
C
O
R
R
E
C
Fig. 1. Development of resting state activity in the rat visual cortex. Spontaneous depth
EEG recorded from layer 4 from three different ages of rat, roughly equivalent to midgestation (P6), third trimester (P10) and term (P13) in human. Direct current recordings
of the field potential reveal infraslow activity dynamics. During the first postnatal week all
activity consists of recurrent infraslow waves that contain rapid (alpha-beta) oscillations
and delta waves, together called a delta brush or spindle burst (P6). These events are
often linked together to form spontaneous activity transients (SATs). Between SATs, cortical networks are largely silent (‘inter-SAT’). SATs continue during the second postnatal
week (P10), but short bursts of activity are now present between SATs. Elimination of retinal activity eliminates SATs, but short activity bursts remain (P10 lower left). Conversely,
pharmacological initiation of constant retinal activity causes constant SAT activity in cortex (P10 lower right). Thus SATs (and therefore 90% of spiking) are not generated spontaneously in cortex but driven by the retina, while the short activity bursts result from true
intrinsic spontaneous activity in thalamocortical circuits. Continuous inter-SAT activity
begins at the end of the second postnatal week (P13), and the size of the now infrequent
SATs is reduced.
Figure modified from Colonnese and Khazipov (2010).
330
Appearance
331
332
In adults, cortical activity patterns are notable for their diversity and
modulation by vigilance state. Some examples are the production of
delta waves of increasing amplitude as the depth of non-REM sleep increases, spindle oscillations during light sleep, alpha oscillations during
eye closure, and gamma-band oscillations during sensory vigilance
(Buzsáki, 2006). Homologues of none of these patterns have been identified in preterm infants. Preterm activity is instead dominated by a few
324
325
326
327
333
334
335
336
337
U
322
323
N
328
329
continuous activity in the adult. The next question is whether the activity bursts that interrupt network silences are fragments of the corticocortical activity that composes the adult resting state, or are of a qualitatively different sort. Along with Pampiglione's ideas quoted above, a
number of facts argue for the latter. In particular (1) the appearance
of the activity is different, (2) the modulation of the cortical activities
by sleep, ‘rest’, and vigilance is poorly developed, (3) adult resting activity does not emerge from immature bursts, and (4) studies in sensory
cortex show that the generative mechanisms are different.
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
Cortical activity in preterm infants is weakly modulated by vigilance
state. Until approximately 32 weeks post-conception both quiet and active sleep (thought to be equivalent to non-REM and REM sleep respectively) are poorly differentiated in EEG and display prominent delta
waves, brushes and SATs as described above. By 32 weeks delta activity
becomes more restricted to quiet sleep and active sleep becomes defined by low amplitude rapid activity, yet as much as 30% of sleep at
these ages is characterized as indeterminate, or containing characteristics of each (Andre et al., 2010). This developmental pattern is matched
in neonatal rats, which display little modulation of cortical activity. Cortical EEG in rats starts to exhibit some sleep/awake state-dependent differentiation of delta activity beginning approximately at P10–11,
becoming robust at the end of the second postnatal week (Frank and
Heller, 1997; Gramsbergen, 1976; Jouvet-Mounier et al., 1970; Seelke
and Blumberg, 2008).
368
369
Developmental separation
383
Simple difference in appearance between the immature and adult
EEG is not sufficient evidence to declare them to be independent phenomena. Functionally similar activity patterns could appear superficially different as a result of changing circuitry or even current source
relationships. In this case we expect that early activity patterns, if followed through development, would transform into adult patterns.
However, in every case examined so far the opposite occurs. The first
studies to identify delta brushes in preterm infants noted their structural similarity to sleep spindles (Dreyfus-Brisac, 1962; Ellingson, 1958;
Watanabe and Iwase, 1972) However, it became rapidly apparent that
delta brushes do not develop into sleep spindles, because sleep spindles
appear a full two months after delta brushes have ceased (Ellingson,
1982). In keeping with these clinical findings, sleep spindles in thalamic
slices from ferrets and mice also appear after the disappearance of delta
brushes (McCormick et al., 1995; Warren and Jones, 1997). A second example is gamma oscillations, which have been proposed to be a
384
Please cite this article as: Colonnese, M., Khazipov, R., Spontaneous activity in developing sensory circuits: Implications for resting state fMRI,
NeuroImage (2012), doi:10.1016/j.neuroimage.2012.02.046
370
371
372
373
374
375
376
377
378
379
380
381
382
385
386
387
388
389
390
391
392 Q9 Q10
393
394
395 Q11
396
397
398
399
M. Colonnese, R. Khazipov / NeuroImage xxx (2012) xxx–xxx
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
C
424
425
E
417
418
R
F
415
416
R
413
414
Peripheral dominance of resting activity during
early development
N
C
O
411
412
Investigators have used the somatosensory and visual cortex of
rats and ferrets to ask the following question: Is resting activity in neonates driven by cortico-cortical connections as it is in adults, or is it
driven by other mechanisms such as spontaneous activity in the sense
organ? These studies have revealed that the large majority of cortical
activity during the period of discontinuity results from input generated spontaneously at the sensory periphery, not within cortex. Examination of electrographic signatures of sensory stimulation in both
neonatal rats and human preterm infants has provided evidence
that a similar peripheral generation of resting activity may occur in
the neonates and fetuses.
This conserved peripheral dominance was first identified in rat primary somatosensory cortex where spontaneous muscle twitches are
associated with local, transient bursts of rhythmic, 5–25 Hz activity,
originally termed ‘spindle burst’(Khazipov et al., 2004)(Fig. 2A). Use of
wide-band recording techniques revealed a very large (up to 1 mV)
negative slow potential associated with rapid oscillations (MarcanoReik and Blumberg, 2008; Minlebaev et al., 2009), thus electrographically resembling delta brushes in human preterms. This was confirmed
in a follow up study in preterm infants that showed delta brushes were
triggered in the somatotopically appropriate region at GW28–32
U
409
410
O
The final evidence that early and late resting activities are different is
that they have fundamentally different pacemakers and generators.
One clear example in cortex is gamma oscillations. In young rodents
gamma oscillations take the form of EGOs, and are primarily determined by gamma-rhythmic excitatory input from thalamus. In the
adult, gamma oscillations depend on perisomatic inhibition within the
cortex. (Minlebaev et al., 2011). This difference in generation leads to
significant differences in the physiological roles of EGOs and adult
gamma oscillations. Whereas EGOs enable the vertical synchronization
of the topographically aligned thalamic and cortical neurons necessary
for the development of thalamocortical units, adult gamma oscillations
synchronize larger areas of the cortex horizontally. A detailed understanding of the thalamic and cortical pacemakers of other early resting
activity patterns is not complete. However, as reviewed in the next section, it is clear that the production of the great majority of activity in the
early sensory cortex is dependent on external drive of thalamus; a fact
that stands in contrast to adult resting activity.
407
408
R
O
422
423
406
P
Generative mechanisms
404
405
(Fig. 2B) (Milh et al., 2007; see also Hrbek et al. (1973) for earlier demonstration of similar phenomenon). Thus in both species direct stimulation of the same body regions induces delta brushes in somatic cortex,
suggesting that it is the sensory feedback produced by the twitch
which triggers delta brushes. In rats spindle bursts do persist after sensory deafferentation (e.g., spinal cord transection), although at a much
reduced frequency (Khazipov et al., 2004). These results suggest that
some component of spindle bursts is endogenous to the thalamocortical
circuit. However, under normal, intact circumstances, they are controlled by the external stimuli brought about by the thalamocortical afferents that trigger these oscillations in the somatosensory cortex in a
somatotopic manner.
Spindle bursts are also the dominant pattern in early postnatal rat
visual cortex (Fig. 1). However, spindle bursts in visual cortex are longer
and less frequent than those in somatosensory cortex. This difference is
due largely to the influence of the retina, not differences intrinsic to the
cortical region. Following enucleation, visual cortex strongly resembles
somatosensory cortex after spinal cord transections, producing shortduration network bursts with long inter-burst periods (Colonnese and
Khazipov, 2010; Hanganu et al., 2006; Khazipov et al., 2004). Retinal involvement in spindle bursts is also indicated by evoked responses. Spindle bursts can be evoked by visual stimulation in preterm infants and
neonatal rats once photoreceptor mediated light responses occur in
the retina (Figs. 2C and D) (Colonnese et al., 2010; Ellingson, 1960;
Hrbek and Mares, 1964), or in rats by electrical stimulation of the eye
before this point (Hanganu et al., 2006). The importance of visual induction of human delta brushes (the homolog of spindle bursts) under normal circumstances is unknown, but is likely to be minimal because the
fetus is relatively visually isolated and spindle occurrence does not depend on the presence of fluctuations of environmental illumination or
movement that might change luminance on the retina. Furthermore,
in rats, the retina is insensitive to light during the first postnatal week,
while spindle bursts are prevalent.
Rather than ambient illumination, spontaneous retinal activity,
which propagates as wave fronts and synchronizes retinal activity within
local domains (Wong, 1999) provides the primary drive for spindle
bursts, at least in rats. A role for these retinal waves in driving visual cortex spindle bursts was demonstrated by simultaneous recordings from
the retina and V1 cortex (Hanganu et al., 2006), which revealed a strong
correlation between the occurrence of spindle bursts in visual cortex and
spontaneous retinal waves. Furthermore, pharmacological manipulations to increase retinal wave generation increased spindle-burst activity
(Colonnese and Khazipov, 2010; Hanganu et al., 2006). Perhaps most
tellingly, the temporal structure of spindle bursts in visual cortex is not
constant during development, but undergoes a profound change during
the second postnatal week, as spindle bursts begin to occur in groups
that together form long-lasting (5–10 s) events associated with large
amplitude DC shifts, highly reminiscent of SATs observed in human preterm neonates. The occurrence of V1 SATs at this time is highly regular,
repeating approximately every 60 s. This change exactly parallels
changes in the structure of spontaneous retinal activity (Blankenship et
al., 2009; Demas et al., 2003; Kerschensteiner and Wong, 2008). Multiunit recordings made from the cortex and thalamus of freely moving ferrets during the same developmental period revealed similar long-lasting
macro-burst behavior eliminated by enucleation (Chiu and Weliky,
2001; Weliky and Katz, 1999).
Thus the differences in the initiation, duration, and intra-burst structure of activity between visual and somatosensory cortex, as well as between visual cortex at different ages, appears to result from differences
inherent in the sense organ input. In total, activity in immature cortex
appears to have a two-tiered generative mechanism: short-duration inputs such as muscle twitches, light flashes, electrical stimulation, and
perhaps even endogenous cortical activity in the absence of peripheral
input, trigger spindle bursts/delta brushes. However, elongated activity
such as retinal waves drives similarly elongated spindle bursts or even
their agglomeration into SATs.
D
421
402
403
T
419
420
mechanism for horizontal synchronization and large scale binding necessary for perception and thought in adults (Buzsáki, 2006). Traditionally thought to emerge relatively late in development, recent
observations in the neonatal rats revealed robust local oscillations at
gamma frequency (Minlebaev et al., 2011). However, despite a similar
frequency and potential to be evoked by sensory stimuli as those of the
adult, the early gamma oscillations (EGOs) are transiently expressed.
They abruptly disappear at the end of the first postnatal week, while
the earliest adult gamma oscillations cannot be observed until days
later, thus arguing against their identity as the same class of activity. A
final example is SATs, introduced above. SATs become elongated in duration but reduced in amplitude with gestational age (Colonnese and
Khazipov, 2010; Tolonen et al., 2007), suggesting that the continuous activity of maturity may emerge due to elongated SATs becoming linked
during development. However, close examination of activity using DC
recordings in both humans and rats showed that continuous resting activity actually emerges from the quiet periods between SATs, and not
from their elongation (Fig. 1). We are not aware of any evidence demonstrating transformation of an identified early activity pattern into a mature network activity. The activities of immaturity studied in detail so
far are clearly temporally independent from adult activity.
E
400
401
5
Please cite this article as: Colonnese, M., Khazipov, R., Spontaneous activity in developing sensory circuits: Implications for resting state fMRI,
NeuroImage (2012), doi:10.1016/j.neuroimage.2012.02.046
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481 Q12
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
6
M. Colonnese, R. Khazipov / NeuroImage xxx (2012) xxx–xxx
A
P5Rat Somatosensory
Cortex
Hindlimb S1 EEG
+
+
-
-
0.1 mV
0.2 s
Hindlimb S1 MUA
0.1 mV
Hindlimb Movement
4s
Right Hand Stimulation
Somatosensory
Cortex
-
F
B GW 30
Twitch - evoked
delta-brush
+
100 µV
T3
FP1 FP2
C3 Cz C4
O
0.5 s
T4
Depth EEG
+
-
*
Sleep
C
+
-
T4
40 µV
Evoked 200 ms
delta-brush
Electrode O2
+
> 2Hz
T3
O1
Cz
500 mV
Evoked
delta-brush
100 mV
500 ms
5s
O
R
O1 O2
MUA
EEG (0.05-40 Hz)
R
FP1 FP2
C3 Cz C4
FP2
C4
T4
O2
FP1
C3
+
400 µV
E
GW 30 Visual
Cortex
T
light flash
T3
Depth EEG
D
400 µV
200 ms
light flash
D
Awake
Evoked
delta-brush
P
P10Rat Visual
Cortex
E
C
R
O
O1 O2
+
528
529
530
531
532
533
534
535
536
537
538
539
540
U
N
C
Fig. 2. Conserved drive of early network oscillations by sensory periphery. In the human and rat somatosensory and visual systems, spontaneous early network oscillations can also be
reliably evoked in a topographically restricted manner by sensory input. A. Simultaneous recording in hindlimb S1 while monitoring hindlimb motion. Spontaneous delta brushes in rats
are associated with spontaneous myclonic jerks, which precede the delta brush. Expanded time scale (right) shows rapid oscillations of delta brushes and associated unit activity following
movements. Direct stimulation of the hind-paw also elicits delta brushes (not shown). B. Direct stimulation of the hand in a preterm infant produces delta brushes in contralateral central
cortex (center panel). Expanded time scale (right) shows delta brush (negative is up) restricted to single electrode. C. Visual stimulation produces delta brushes in visual cortex. Responses
to 6 light flashes in layer 4 of V1 from the same animal during different vigilance states (central traces). A spontaneous delta brush occurs before one of the stimulus trials (asterisk). Note
the similarity of spontaneous and evoked patterns. Expanded time scale (right) shows field potential and evoked unit activity. D. Visual stimulation produces delta brushes in occipital
electrodes of preterm infants (center panel). Expanded time scale (right) shows evoked potentials similar to the rodent visual response.
Modified from Khazipov et al. (2004); Milh et al. (2007); and Colonnese et al. (2010).
Recent studies suggest that regions outside of the primary sensory
regions follow similar developmental patterns. Prefrontal cortex demonstrates alternating periods of silence and rapid oscillation that are
driven by the hippocampus and become continuous around the same
time as sensory cortex (Brockmann et al., 2011). Similar activity patterns have also been described in dysgranular and temporal cortex
(Adelsberger et al., 2005; Seelke and Blumberg, 2010), though their
generative mechanisms are still unclear.
In summary, recent rodent work, partially confirmed in human preterm infants, shows that the resting state neural events (SATs/spindle
bursts/delta brushes) which constitute the majority of early neural activity (>90% in rat visual cortex) have as their primary driver input
from the peripheral sense organ (Fig. 3). In adults the proportion of
resting activity contributed by sensory input to primary sensory cortex
is likely to be small, as sensory input is specifically kept minimal during
testing and ongoing activity in visual cortex is only minimally modulated by visual inputs (Fiser et al., 2004).
541
Development of intrinsic thalamocortical activities
545
If the basis of resting state activity modulations in sensory cortex during early development is spontaneous activation at the periphery, then
they result from different functional circuits from those generating resting activity in the adult. The activity patterns most closely associated
with resting state modulations in sensory cortex are delta waves (Jann
et al., 2010; Lu et al., 2007), which result from spontaneous activation
546
Please cite this article as: Colonnese, M., Khazipov, R., Spontaneous activity in developing sensory circuits: Implications for resting state fMRI,
NeuroImage (2012), doi:10.1016/j.neuroimage.2012.02.046
542
543
544
547
548
549
550
551
M. Colonnese, R. Khazipov / NeuroImage xxx (2012) xxx–xxx
Immature
Pre-resting state
7
Adult
Resting state
Higher SC
and other Ctx
Higher
VC and
other Ctx
Higher
SC and
other Ctx
Visual
Cortex
Somatosensory
Cortex
Visual
Cortex
SC
Thalamus
Thalamus
Thalamus
Thalamus
Retina
(spontaneous
waves)
Somatic
Receptors
(spontaneous
twitches)
Retina
F
HigherVC
and other Ctx
O
Somatic
Receptors
564
565
566
567
568
569
570
Q13571
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
D
in rodents (Gramsbergen, 1976). This rapid increase in spontaneous activity does not emerge simultaneously in entire cortex, but instead occurs first in somatosensory and then in visual cortex in both the rat
and human. In both species it coincides with the elimination of delta
brushes and spindle bursts in these areas (Andre et al., 2010).
At least two developmental factors relevant to the generation of
resting state connectivity are likely involved in the emergence of
the activated cortical state: the development of arousal systems and
intra-cortical connections. Isolating the forebrain from ascending
neuromodulatory arousal inputs at P13–15 rats reverses the developmental increase in spontaneous cortical activity, reinstating the immature discontinuity (Colonnese et al., 2010). The mature activity
state is re–instated by application of norepinephrine, a key initiator
of the activated cortical state (Constantinople and Bruno, 2011).
This is consistent with the critical role for noradrenergic systems for
the development of sleep-wake transitions (Gall et al., 2009) and
the maturation of cholinergic (Mechawar and Descarries, 2001;
Robertson et al., 1985) and noradrenergic afferents and receptor distributions (Latsari et al., 2002; Venkatesan et al., 1996) around this
time. However, arousal system activity alone is incapable of inducing
a continuous active state in the immature cortex (Hanganu et al.,
2007). Thus the acquisition of brainstem control of cortical rest states
likely requires thalamocortical inputs as well. A prominent candidate
is the strengthening of the long and short range intra-cortical connectivity, required for maintenance of the activated cortical state
(Sanchez-Vives and McCormick, 2000; Steriade and McCarley, 2005;
Timofeev et al., 2000). The network of horizontal connections between cortical columns emerges shortly before eye opening in cats
and ferrets (Durack and Katz, 1996; Ruthazer and Stryker, 1996) independently of visual input (Callaway and Katz, 1990; Durack and
Katz, 1996; Ruthazer and Stryker, 1996). In mice feedback connections to V1 are observed just before eye opening (Berezovskii et al.,
2011). In humans, dense horizontal connections are also first observed at GW37 (Burkhalter et al., 1993). Emergence of the continuous mode of cortical activity thus likely results from a coincidence
of two developmentally regulated factors: (i) maturation of the brainstem arousal input to thalamus and cortex, which provides tonic neuronal depolarization and (ii) formation of excitatory synaptic
connections among local and distant cortical neurons, required for
initiation and maintenance of the activated cortical state, and for resting state modulations.
E
T
C
562
563
E
560
561
R
558
559
R
556
557
N
C
O
554
555
of thalamocortical circuits during sleep or periods of low vigilance
(Steriade and McCarley, 2005). Not surprisingly, the development of
delta activity induced by non-REM sleep parallels the loss of early activity
patterns in rats and humans (Andre et al., 2010; Gramsbergen, 1976;
Seelke and Blumberg, 2008). Thus, the question of when adult resting
state develops a question of when intrinsic activity generated within cortical circuits develops (Fig. 3). In adults, intrinsically generated activity in
thalamocortical circuits is characterized by the presence of activated cortical states (aka desynchronized- or up-state) (Destexhe et al., 2003,
2007). This intrinsic activity is generated via recurrent cortical connectivity, which drives balanced excitatory and inhibitory input to neurons,
and is regulated by neuromodulatory input, particularly acetylcholine
and norephinephrine (Steriade and McCarley, 2005). The active state occurs during waking, REM sleep and even slow wave sleep, when it is broken up by very short periods of network silence, or down states.
Intracellular recordings show little difference between continuous cortical activity during awake, REM sleep and UP-states during slow wave
sleep. As described in Sections Discontinuous temporal organization of
early activity, Unique activity patterns in preterm infants, and
Peripheral dominance of resting activity during early development, the
current evidence suggests that the activated cortical state is not present
in neonatal rats and young premature neonates, as the majority of activity is driven directly by the periphery, not intra-cortical circuits. Furthermore, the early sensory driven SAT and delta brushes do not develop
into activated cortical states. Instead it appears that primitive active
states, which resemble those observed in isolated cortical slabs, develop
independently during the second postnatal week in the rat, which show
increased delta activity during quiet sleep (Seelke and Blumberg, 2008),
calcium bursts under anesthesia (Golshani et al., 2009; Rochefort et al.,
2009) and short 200–500 ms bursts of cortical activity that are insensitive to eye-removal (Colonnese and Khazipov, 2010). This may correspond to the increase in delta activity observed during quiet sleep
during third trimester preterm infants, or the tracé alternant pattern
which becomes increasingly prevalent during active sleep toward the
end of gestation. Interestingly, rat cortical slices begin to produce spontaneous synaptic activity around the same time (Allene et al., 2008;
Rheims et al., 2008), demonstrating a basic change in the ability of isolated cortex to generate activity. In visual cortex these early delta/slow
waves become rapidly more frequent around eye opening (Colonnese
and Khazipov, 2010; Colonnese et al., 2010; Rochefort et al., 2009), coinciding with the emergence of clearly differentiable EEG sleep states
U
552
553
P
R
O
Fig. 3. Changing drive of resting state activity during development of sensory systems. Diagram shows hypothesized circuits influencing cortical activity during preterm and adult
resting states in primary sensory systems. During early development in the absence of visual input spontaneous waves cause the retina to be synchronously active. Spontaneous
jerks serve a similar purpose in the somatosensory system. During this period spontaneous intra-cortical activity (both local and feedback) is minimal. The majority of resting activity thus results from the transmission of spontaneous retinal activity. This is in contrast to adults, whose resting activity is largely independent of retinal or somatic input and is
instead generated via recurrent interactions between and within cortex and thalamus. Because of these substantial differences in drive and function between the immature and
adult systems, we propose that the preterm infant and neonatal rat be considered to be in a pre-resting state.
Please cite this article as: Colonnese, M., Khazipov, R., Spontaneous activity in developing sensory circuits: Implications for resting state fMRI,
NeuroImage (2012), doi:10.1016/j.neuroimage.2012.02.046
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
C
660
661
E
658
659
R
656
657
R
654
655
O
652
653
C
650
651
N
648
649
U
646
647
Summary
744
Electrophysiological evidence in human preterm infants and animal
models reviewed here is consistent with the hypothesis that early circuit development is characterized by a unique state of cortical network
dynamics, in which adult-like thalamocortical resting state activity is
minimal, allowing inputs from the periphery to be amplified by cortical
oscillatory bursts (Fig. 3). This period appears tightly regulated to coincide with the sensory deprivation experienced in the womb (or maintained by eye or ear canal closure in rodents), and to end when active
sensory processing is initiated. It is only when this early period ends
that many of the gross characteristics of adult resting state activity can
be discerned, including clear regulation of cortical EEG by sleep state
and the reduction and eventual elimination of SATs in counterpoint to
continuous activity (Gramsbergen, 1976; Seelke and Blumberg, 2008;
Vanhatalo et al., 2005), and perhaps not surprisingly, elements of complex cortico-cortical resting state networks (Doria et al., 2010; Fransson
et al., 2011; Gao et al., 2009). Of course, the development of cortical activity is by no means complete by this point; in essence it has only just
745
746
F
We have presented evidence that ‘resting state’ modulations of activity in sensory cortex during early ages are not the same infraslow
modulation present in the adult. This is supported by two characteristics of neocortical activity during the fetal and early neonatal period
(1) unique early network activities driven by the sensory periphery
and (2) the lack of intrinsic activity generated in thalamocortical circuits. The young resting state activities represent a unique network
state that is limited to early development. We propose this state is optimized to drive synaptic plasticity in thalamocortical networks in response to spontaneous activity generated in the sense organs. To
support this function, ‘resting’ activity at these ages largely reflects activity generated in the sensory periphery, despite its independence
from changes in the external environment. As a result, cortical activity
at these early ages is poorly modulated by vigilance state or the environment; in effect the cortex is always ‘at rest’ in the sense that there
are no task or attention based modulations of the cortex. Unlike adults,
however, rest is a singular condition, at least during early development.
A key implication of this hypothesis for imaging studies is that while
powerful activity modulations are present during early development,
they are incapable of conveying graded and high-frequency information
transfer as in adults.
Two examples support this assertion. Using two-photon calcium imaging of the superficial layers of visual cortex in anesthetized mice
Rochefort et al. (2009) and Golshani et al. (2009) traced the development of cortical slow waves. They observed a rapid developmental decrease in neuronal synchrony during individual slow waves by the end
of the second postnatal week. Why is this important? Generation of active states is proposed to be a primary mode of intra-cortical communication, by activation of particular circuits and not others during waking
and by driving replay of memories during sleep (Destexhe et al., 2003; Ji
and Wilson, 2007), both of which require differential firing of neurons
based on behavioral or sensory experience. Thus synchronous firing
by all neurons is poorly adapted for information processing, though it
is well adapted for the stabilization of grossly refined connectivity during early development.
Such all or none responses are found during early sensory processing. Colonnese et al. (2010) examined the role of early thalamocortical
networks properties in determining visual responses in preterm
humans and rat. Despite weak or absent retinal visual responses in
vitro, robust light responses could be evoked as early as P8 in rats. As
discussed previously, cortical light responses consisted of a complex
group of network oscillations including delta brushes (Fig. 2C). These
visual responses were ‘all-or-none’ bursts, effectively registering the
presence of an even minimal stimulation, but not its intensity. Measurement of the spiking responses of retinal ganglion cells in acutely excised
retina allowed further quantification of the input-output relationships,
which showed that these early bursting responses massively amplified
retinal activity within the cortical circuit. Only with the development of
continuous activity in cortex were graded responses to light observed.
Thus we suggest that the primary role of early activity patterns, in addition to precisely synchronizing activity, is to amplify weak, tentative inputs. Such amplification is incompatible with visual processing. These
changes in bursting and amplification observed in the visual system
641
642
O
645
640
R
O
Early cortical function and implications for resting state fMRI
638
639
698
699
P
644
636
637
(where they can be quantified) are likely also to occur in other, nonsensory, cortical areas, where we would predict that early activity
might be similarly ’non-coding’, but instead serve more limited synapse
stabilization goals at these early ages.
In short these data strongly argue that while correlated activity within and between various brain regions is undoubtedly present during the
fetal period, they are much more likely to reflect the presence of anatomical connections rather than a functional transfer of information. In fact
two pieces of evidence suggest that early connectivity studies may
‘over-estimate’ connectivity in preterm infants relative to adult. First,
the lack of continuous activity should increase signal to noise, leading
to detection of connectivity levels below those apparent in adults. Second, the amplification of weak inputs should result in activity levels
with much lower anatomical connectivity indexes.
By contrast, fMRI has a potential to underestimate connectivity in
young brains relative to electrophysiological measurements because
neurovascular coupling is still immature. While relatively tight coupling
of neuronal activity and the Blood Oxygen Level Dependent (BOLD) signal in adults has been established in animal models (Logothetis and
Wandell, 2004), the developmental timing of this coupling is poorly explored. In rat somatosensory cortex, evoked BOLD signal in anesthetized
rats could not be observed before the 13th postnatal day (Colonnese et
al., 2008). Even after the development of functional coupling, the time
course of hemodynamic responses evolved over an extended time period, which should change the relevant frequencies of resting state modulations. Combined studies of spontaneous cortical EEG and blood flow
measured with near infrared spectroscopy in preterm infants show that
SATs are accompanied by increased oxygen extraction leading to a large
and extended decrease in oxy-hemoglobin which is followed by a
delayed hyperfusion of oxygenated blood relative to the adult value
(Roche-Labarbe et al., 2007), somewhat resembling early BOLD signals
in rats. fMRI studies in term infants have suggested a changing developmental relationship between evoked neural activity and the BOLD signal. Pattern vision evoked BOLD signals undergo an inversion, moving
from positive to negative around 8 weeks (Muramoto et al., 2002).
The authors suggest this is a result of rapid synaptogenesis out-pacing
vascular development at these ages, causing oxygen extraction to exceed reactive blood flow. This effect has seen mixed support in other
studies (Anderson et al., 2001; Dehaene-Lambertz et al., 2002; Erberich
et al., 2003). Thus until the development of the neurovascular unit is
better understood in humans and animal models, methods relying on
neurovascular dynamics in the infant, particularly preterm infants, do
not have the same experimental support as currently exists for adults,
and have the potential to underestimate connectivity, while the activity
and electrical characteristics of young tissue may provide a counter balancing amplification.
T
643
In summary, spontaneous activity generated within thalamocortical
circuits, i.e. the true ‘resting state’, is developmentally delayed relative
to cortical responses to spontaneous activity in the periphery, which
generate transient excitatory waves on an infraslow timescale during
early development. Only after the establishment of basic cortical circuits, such as topographic mapping of thalamic inputs, does significant
intra-cortical connectivity and modulation by vigilance state emerge.
It is perhaps then not surprising that resting state networks involving
multiple, distant cortical regions emerge after local, primarily sensory,
networks (Doria et al., 2010; Power et al., 2010).
D
634
635
M. Colonnese, R. Khazipov / NeuroImage xxx (2012) xxx–xxx
E
8
Please cite this article as: Colonnese, M., Khazipov, R., Spontaneous activity in developing sensory circuits: Implications for resting state fMRI,
NeuroImage (2012), doi:10.1016/j.neuroimage.2012.02.046
700
701
702
703
704
705
706
707
708
709
710
711
712
713
714
715
716
717
718
719
720
721
722
723
724
725
726
727
728
729
730
731
732
733
734
735
736
737
738
739
740
741
742
743
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
M. Colonnese, R. Khazipov / NeuroImage xxx (2012) xxx–xxx
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
Q15799
801
802
803
804
805
806
807
808
809
810
811
812
813
814
815
816
817
818
819
820
821
822
823
824
825
826
827
828
829
830
831
832
833
F
O
R
O
Q14780
Uncited references
Minlebaev et al., 2007
Yang et al., 2009
P
779
References
Adelsberger, H., Garaschuk, O., Konnerth, A., 2005. Cortical calcium waves in resting
newborn mice. Nat. Neurosci. 8, 988–990.
Allene, C., Cattani, A., Ackman, J.B., Bonifazi, P., Aniksztejn, L., Ben-Ari, Y., Cossart, R.,
2008. Sequential generation of two distinct synapse-driven network patterns in
developing neocortex. J. Neurosci. 28, 12851–12863.
Anderson, A.W., Marois, R., Colson, E.R., Peterson, B.S., Duncan, C.C., Ehrenkranz, R.A.,
Schneider, K.C., Gore, J.C., Ment, L.R., 2001. Neonatal auditory activation detected
by functional magnetic resonance imaging. Magn. Reson. Imaging 19, 1–5.
Andre, M., Lamblin, M.D., d'Allest, A.M., Curzi-Dascalova, L., Moussalli-Salefranque, F., T,
S.N.T., Vecchierini-Blineau, M.F., Wallois, F., Walls-Esquivel, E., Plouin, P., 2010.
Electroencephalography in premature and full-term infants. Developmental features and glossary. Neurophysiol. Clin. 40, 59–124.
Baria, A.T., Baliki, M.N., Parrish, T., Apkarian, A.V., 2011. Anatomical and functional assemblies of brain BOLD oscillations. J. Neurosci. 31, 7910–7919.
Berezovskii, V.K., Nassi, J.J., Born, R.T., 2011. Segregation of feedforward and feedback
projections in mouse visual cortex. J. Comp. Neurol.
Bergstrom, R.M., Hellstrom, P.E., Stenberg, D., 1962. On the development of electrical
activity in the cerebrum of the intrauterine guinea pig fetus. Ann. Chir. Gynaecol.
Fenn. 51, 466–474.
Berkes, P., Orban, G., Lengyel, M., Fiser, J., 2011. Spontaneous cortical activity reveals
hallmarks of an optimal internal model of the environment. Science 331, 83–87.
Bernhard, C.G., Kaiser, I.H., Kolmodin, G.M., 1959. On the development of cortical activity in fetal sheep. Acta Physiol. Scand. 47, 333–349.
Bianciardi, M., Fukunaga, M., van Gelderen, P., Horovitz, S.G., de Zwart, J.A., Duyn, J.H.,
2009. Modulation of spontaneous fMRI activity in human visual cortex by behavioral state. Neuroimage 45, 160–168.
Blankenship, A.G., Feller, M.B., 2010. Mechanisms underlying spontaneous patterned
activity in developing neural circuits. Nat. Rev. Neurosci. 11, 18–29.
Blankenship, A.G., Ford, K.J., Johnson, J., Seal, R.P., Edwards, R.H., Copenhagen, D.R.,
Feller, M.B., 2009. Synaptic and extrasynaptic factors governing glutamatergic retinal waves. Neuron 62, 230–241.
Brockmann, M.D., Poschel, B., Cichon, N., Hangau-Opatz, I.L., 2011. Coupled oscillations
mediated directed interaction between prefrontal cortex and hippocampus of the
neonatal rat. Neuron 71, 332–347.
Burkhalter, A., Bernardo, K.L., Charles, V., 1993. Development of local circuits in human
visual cortex. J. Neurosci. 13, 1916–1931.
Buzsáki, G., 2006. Rhythms of the Brain. Oxford University Press, Oxford ; New York.
Callaway, E.M., Katz, L.C., 1990. Emergence and refinement of clustered horizontal connections in cat striate cortex. J. Neurosci. 10, 1134–1153.
Charles, M.S., Fuller, J.L., 1956. Developmental study of the electroencephalogram of
the dog. Electroencephalogr. Clin. Neurophysiol. 8, 645–652.
Chiu, C., Weliky, M., 2001. Spontaneous activity in developing ferret visual cortex in
vivo. J. Neurosci. 21, 8906–8914.
Clancy, B., Darlington, R.B., Finlay, B.L., 2000. The course of human events: predicting
the timing of primate neural development. Dev. Sci. 3, 57–66.
Clancy, B., Darlington, R.B., Finlay, B.L., 2001. Translating developmental time across
mammalian species. Neuroscience 105, 7–17.
Clancy, B., Kersh, B., Hyde, J., Darlington, R.B., Anand, K.J.S., Finlay, B.L., 2007. Webbased method for translating neurodevelopment from laboratory species to
humans. Neuroinformatics 5, 79–94.
D
777
778
Colonnese, M.T., Khazipov, R., 2010. "Slow activity transients" in infant rat visual cortex: a spreading synchronous oscillation patterned by retinal waves. J. Neurosci.
30, 4325–4337.
Colonnese, M.T., Phillips, M.A., Constantine-Paton, M., Kaila, K., Jasanoff, A., 2008. Development of hemodynamic responses and functional connectivity in rat somatosensory cortex. Nat. Neurosci. 11, 72–79.
Colonnese, M.T., Kaminska, A., Minlebaev, M., Milh, M., Bloem, B., Lescure, S., Moriette,
G., Chiron, C., Ben-Ari, Y., Khazipov, R., 2010. A conserved switch in sensory processing prepares developing neocortex for vision. Neuron 67, 480–498.
Constantinople, C.M., Bruno, R.M., 2011. Effects and mechanisms of wakefulness on
local cortical networks. Neuron 69, 1061–1068.
Contreras, D., Timofeev, I., Steriade, M., 1996. Mechanisms of long-lasting hyperpolarizations underlying slow sleep oscillations in cat corticothalamic networks. J. Physiol. 494 (Pt 1), 251–264.
Crain, S.M., 1952. Development of electrical activity in the cerebral cortex of the albino
rat. Proc. Soc. Exp. Biol. Med. 81, 49–51.
Curzi-Dascalova, L., 1977. Waking and sleeping E.E.G. in normal babies before 6 months
of age (author's transl). Rev. Electroencephalogr. Neurophysiol. Clin. 7, 316–326.
Dehaene-Lambertz, G., Dehaene, S., Hertz-Pannier, L., 2002. Functional neuroimaging
of speech perception in infants. Science 298, 2013–2015.
Demas, J., Eglen, S.J., Wong, R.O., 2003. Developmental loss of synchronous spontaneous activity in the mouse retina is independent of visual experience. J. Neurosci.
23, 2851–2860.
Destexhe, A., Rudolph, M., Paré, D., 2003. The high-conductance state of neocortical
neurons in vivo. Nat. Rev. Neurosci. 4, 739–751.
Destexhe, A., Hughes, S.W., Rudolph, M., Crunelli, V., 2007. Are corticothalamic 'up'
states fragments of wakefulness? Trends Neurosci. 30, 334–342.
Deza, L., Eidelberg, E., 1967. Development of cortical electrical activity in the rat. Exp.
Neurol. 17, 425–438.
Doria, V., Beckmann, C.F., Arichi, T., Merchant, N., Groppo, M., Turkheimer, F.E.,
Counsell, S.J., Murgasova, M., Aljabar, P., Nunes, R.G., Larkman, D.J., Rees, G.,
Edwards, A.D., 2010. Emergence of resting state networks in the preterm human
brain. Proc. Natl. Acad. Sci. 107, 20015–20020.
Dreyfus-Brisac, C., Larroche, J.C., 1971. Discontinuous electroencephalograms in the
premature newborn and at term. Electro-anatomo-clinical correlations. Rev. Electroencephalogr. Neurophysiol. Clin. 1, 95–99.
Dreyfus-Brisac, C., Monod, N., 1965. Sleep of premature and full-term neonates—a
polygraphic study. Proc. R. Soc. Med. 58, 6–7.
Dreyfus-Brisac, C., Fischgold, H., Samson-Dollfus, D., Saint-Anne Dargassies, S., Ziegler,
T., Monod, N., Blanc, C., 1956. Veille sommeil et reactivite sensorielle chez le premature et le nouveau-ne. Electroencephalogr. Clin. Neurophysiol. 6, 418–440.
Durack, J.C., Katz, L.C., 1996. Development of horizontal projections in layer 2/3 of ferret visual cortex. Cereb. Cortex 6, 178–183.
Ellingson, R.J., 1960. Cortical electrical responses to visual stimulation in the human infant. Electroencephalogr. Clin. Neurophysiol. 12, 663–677.
Ellingson, R.J., 1970. Ontogenesis of the electroencephalogram. In: Himwich, H.A. (Ed.),
Developmental Neurobiology. Charles C Thomas, pp. 441–474.
Ellingson, R.J., Peters, J.F., 1980. Development of EEG and daytime sleep patterns in low
risk premature infants during the first year of life: longitudinal observations. Electroencephalogr. Clin. Neurophysiol. 50, 165–171.
Erberich, S.G., Friedlich, P., Seri, I., Nelson Jr., M.D., Blum, S., 2003. Functional MRI in neonates
using neonatal head coil and MR compatible incubator. Neuroimage 20, 683–692.
Eswaran, H., Lowery, C.L., Wilson, J.D., Murphy, P., Preissl, H., 2004. Functional development of the visual system in human fetus using magnetoencephalography. Exp.
Neurol. 190 (Suppl. 1), S52–S58.
Eswaran, H., Haddad, N.I., Shihabuddin, B.S., Preissl, H., Siegel, E.R., Murphy, P., Lowery,
C.L., 2007. Non-invasive detection and identification of brain activity patterns in
the developing fetus. Clin. Neurophysiol. 118, 1940–1946.
Fiser, J., Chiu, C., Weliky, M., 2004. Small modulation of ongoing cortical dynamics by
sensory input during natural vision. Nature 431, 573–578.
Fox, M.W., 1967. Postnatal development of the EEG in the dog. 3. Summary and discussion of development of canine EEG. J. Small Anim. Pract. 8, 109–112.
Fox, M.D., Raichle, M.E., 2007. Spontaneous fluctuations in brain activity observed with
functional magnetic resonance imaging. Nat. Rev. Neurosci. 8, 700–711.
Frank, M.G., Heller, H.C., 1997. Development of diurnal organization of EEG slow-wave
activity and slow-wave sleep in the rat. Am. J. Physiol. 273, R472–R478.
Fransson, P., Skiold, B., Horsch, S., Nordell, A., Blennow, M., Lagercrantz, H., Aden, U., 2007.
Resting-state networks in the infant brain. Proc. Natl. Acad. Sci. 104, 15531–15536.
Fransson, P., Aden, U., Blennow, M., Lagercrantz, H., 2011. The functional architecture of
the infant brain as revealed by resting-state FMRI. Cereb. Cortex 21, 145–154.
Gall, A.J., Joshi, B., Best, J., Florang, V.R., Doorn, J.A., Blumberg, M.S., 2009. Developmental emergence of power-law wake behavior depends upon the functional integrity
of the locus coeruleus. Sleep 32, 920–926.
Gao, W., Zhu, H., Giovanello, K.S., Smith, J.K., Shen, D., Gilmore, J.H., Lin, W., 2009. Evidence on the emergence of the brain's default network from 2-week-old to 2-yearold healthy pediatric subjects. Proc. Natl. Acad. Sci. 106, 6790–6795.
Golshani, P., Gonsalves, J.T., Khoshkhoo, S., Mostany, R., Smirnakis, S., Portera-Cailliau,
C., 2009. Internally mediated developmental desynchronization of neocortical network activity. J. Neurosci. 29, 10890–10899.
Gramsbergen, A., 1976. The development of the EEG in the rat. Dev. Psychobiol. 9, 501–515.
Guzzetta, A., Baldini, S., Bancale, A., Baroncelli, L., Ciucci, F., Ghirri, P., Putignano, E., Sale,
A., Viegi, A., Berardi, N., Boldrini, A., Cioni, G., Maffei, L., 2009. Massage accelerates
brain development and the maturation of visual function. J. Neurosci. 29,
6042–6051.
Haider, B., McCormick, D.A., 2009. Rapid neocortical dynamics: cellular and network
mechanisms. Neuron 62, 171–189.
E
775
776
T
773
774
C
771
772
E
769
770
R
768
R
766
767
N
C
O
764
765
begun (Power et al., 2010). For example, cortical activity in ferrets presented with various visual stimuli just after eye opening is still
dominated by highly-correlated spontaneous activity, and it is not
until the critical period for monocular deprivation that cortical activity
has become dominated by internal dynamics (Berkes et al., 2011).
Unfortunately, our current knowledge is focused on primary sensory systems, leaving open the possibility that development in association areas most commonly ‘activated’ during the resting state follow
different trajectories. Furthermore, mechanistic and invasive studies
have focused on rodents, which provide a poor model for the massively expanded and folded human brain. Nevertheless, the developmental changes in the origins, transmission, and function of activity
in sensory cortex reviewed here should be kept in mind for the interpretation of imaging studies during these time periods. Ideally, multimodal studies using electrophysiological measures with imaging
combined with carefully designed animal studies will reveal the developmental mechanisms of higher cortical functions such as the development of default state and attentional networks.
U
762
763
9
Please cite this article as: Colonnese, M., Khazipov, R., Spontaneous activity in developing sensory circuits: Implications for resting state fMRI,
NeuroImage (2012), doi:10.1016/j.neuroimage.2012.02.046
834
835
836
837
838
839
840
841
842
843
844
845
846
847
848
849
850
851
852
853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
881
882
883
884
885
886
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
902
903
904
905
906
907
908
909
910
911
912
913
914
915
916
917
918
919
D
P
R
O
O
F
Power, J.D., Fair, D.A., Schlaggar, B.L., Petersen, S.E., 2010. The development of human
functional brain networks. Neuron 67, 735–748.
Rheims, S., Minlebaev, M., Ivanov, A., Represa, A., Khazipov, R., Holmes, G.L., Ben-Ari, Y.,
Zilberter, Y., 2008. Excitatory GABA in rodent developing neocortex in vitro. J. Neurophysiol. 100, 609–619.
Robertson, R.T., Tijerina, A.A., Gallivan, M.E., 1985. Transient patterns of acetylcholinesterase activity in visual cortex of the rat: normal development and the effects of
neonatal monocular enucleation. Brain Res. 353, 203–214.
Rochefort, N.L., Garaschuk, O., Milos, R.-I., Narushima, M., Marandi, N., Pichler, B.,
Kovalchuk, Y., Konnerth, A., 2009. Sparsification of neuronal activity in the visual
cortex at eye-opening. Proc. Natl. Acad. Sci. 106, 15049–15054.
Roche-Labarbe, N., Wallois, F., Ponchel, E., Kongolo, G., Grebe, R., 2007. Coupled oxygenation oscillation measured by NIRS and intermittent cerebral activation on
EEG in premature infants. Neuroimage 36, 718–727.
Rose, G.H., Ellingson, R.J., 1968. The comparative ontogenesis of visually evoked responses
in rat, cat and human infant. Electroencephalogr. Clin. Neurophysiol. 24, 284–285.
Ruthazer, E.S., Stryker, M.P., 1996. The role of activity in the development of long-range
horizontal connections in area 17 of the ferret. J. Neurosci. 16, 7253–7269.
Sadaghiani, S., Hesselmann, G., Friston, K.J., Kleinschmidt, A., 2010. The relation of ongoing
brain activity, evoked neural responses, and cognition. Front. Syst. Neurosci. 4, 20.
Sanchez-Vives, M.V., McCormick, D.A., 2000. Cellular and network mechanisms of
rhythmic recurrent activity in neocortex. Nat. Neurosci. 3, 1027–1034.
Scherrer, J., Oeconomos, D., 1954. Comparison of somatesthetic cortical responses in
newborn and adult animals. Etudes Neonatales 3, 199–216.
Seelke, A.M., Blumberg, M.S., 2008. The microstructure of active and quiet sleep as cortical delta activity emerges in infant rats. Sleep 31, 691–699.
Seelke, A.M., Blumberg, M.S., 2010. Developmental appearance and disappearance of
cortical events and oscillations in infant rats. Brain Res. 1324, 34–42.
Shu, Y., Hasenstaub, A., McCormick, D.A., 2003. Turning on and off recurrent balanced
cortical activity. Nature 423, 288–293.
Smyser, C.D., Inder, T.E., Shimony, J.S., Hill, J.E., Degnan, A.J., Snyder, A.Z., Neil, J.J., 2010.
Longitudinal analysis of neural network development in preterm infants. Cereb.
Cortex 20, 2852–2862.
Steriade, M., McCarley, R.W., 2005. Brain Control of Wakefulness and Sleep. Springer,
New York, N.Y.
Steriade, M., Nunez, A., Amzica, F., 1993. A novel slow (b 1 Hz) oscillation of neocortical
neurons in vivo: depolarizing and hyperpolarizing components. J. Neurosci. 13,
3252–3265.
Timofeev, I., Grenier, F., Bazhenov, M., Sejnowski, T.J., Steriade, M., 2000. Origin of slow
cortical oscillations in deafferented cortical slabs. Cereb. Cortex 10, 1185–1199.
Tolonen, M., Palva, J.M., Andersson, S., Vanhatalo, S., 2007. Development of the spontaneous activity transients and ongoing cortical activity in human preterm babies.
Neuroscience 145, 997–1006.
Tritsch, N.X., Yi, E., Gale, J.E., Glowatzki, E., Bergles, D.E., 2007. The origin of spontaneous activity in the developing auditory system. Nature 450, 50–55.
Tuge, H., Kanayama, Y., Chang, H.Y., 1960. Comparative studies on the development of
EEG. Jpn. J. Physiol. 10, 211–220.
Vanhatalo, S., Kaila, K., 2006. Development of neonatal EEG activity: from phenomenology to physiology. Semin. Fetal Neonatal Med. 11, 471–478.
Vanhatalo, S., Tallgren, P., Andersson, S., Sainio, K., Voipio, J., Kaila, K., 2002. DC-EEG discloses prominent, very slow activity patterns during sleep in preterm infants. Clin.
Neurophysiol. 113, 1822–1825.
Vanhatalo, S., Palva, J.M., Andersson, S., Rivera, C., Voipio, J., Kaila, K., 2005. Slow endogenous activity transients and developmental expression of K+–Cl − cotransporter
2 in the immature human cortex. Eur. J. Neurosci. 22, 2799–2804.
Venkatesan, C., Song, X.Z., Go, C.G., Kurose, H., Aoki, C., 1996. Cellular and subcellular
distribution of alpha 2A-adrenergic receptors in the visual cortex of neonatal and
adult rats. J. Comp. Neurol. 365, 79–95.
Vincent, J.L., Patel, G.H., Fox, M.D., Snyder, A.Z., Baker, J.T., Van Essen, D.C., Zempel, J.M.,
Snyder, L.H., Corbetta, M., Raichle, M.E., 2007. Intrinsic functional architecture in
the anaesthetized monkey brain. Nature 447, 83–86.
Warren, R.A., Jones, E.G., 1997. Maturation of neuronal form and function in a mouse
thalamo-cortical circuit. J. Neurosci. 17, 277–295.
Watanabe, K., Iwase, K., 1972. Spindle-like fast rhythms in the EEGs of low-birth
weight infants. Dev. Med. Child Neurol. 14, 373–381.
Weliky, M., Katz, L.C., 1999. Correlational structure of spontaneous neuronal activity in
the developing lateral geniculate nucleus in vivo. Science 285, 599–604.
Wong, R.O., 1999. Retinal waves and visual system development. Annu. Rev. Neurosci.
22, 29–47.
Yang, J.-W., Hanganu-Opatz, I.L., Sun, J.-J., Luhmann, H.J., 2009. Three patterns of oscillatory activity differentially synchronize developing neocortical networks in vivo. J.
Neurosci. 29, 9011–9025.
Yoshii, N., Tsukiyama, K., 1951. Normal EEG and its development in the white rat. Jpn. J.
Physiol. 2, 34–38.
Zuo, X.N., Di Martino, A., Kelly, C., Shehzad, Z.E., Gee, D.G., Klein, D.F., Castellanos, F.X.,
Biswal, B.B., Milham, M.P., 2010. The oscillating brain: complex and reliable. Neuroimage 49, 1432–1445.
N
C
O
R
R
E
C
T
Hanganu, I.L., Ben-Ari, Y., Khazipov, R., 2006. Retinal waves trigger spindle bursts in the
neonatal rat visual cortex. J. Neurosci. 26, 6728–6736.
Hanganu, I.L., Staiger, J.F., Ben-Ari, Y., Khazipov, R., 2007. Cholinergic modulation of
spindle bursts in the neonatal rat visual cortex in vivo. J. Neurosci. 27, 5694–5705.
Hasenstaub, A., Sachdev, R.N.S., McCormick, D.A., 2007. State changes rapidly modulate
cortical neuronal responsiveness. J. Neurosci. 27, 9607–9622.
Hrbek, A., Mares, P., 1964. Cortical evoked responses to visual stimulation in full-term
and premature newborns. Electroencephalogr. Clin. Neurophysiol. 16, 575–581.
Hrbek, A., Karlberg, P., Olsson, T., 1973. Development of visual and somatosensory
evoked responses in pre-term newborn infants. Electroencephalogr. Clin. Neurophysiol. 34, 225–232.
Huberman, A.D., Feller, M.B., Chapman, B., 2008. Mechanisms underlying development
of visual maps and receptive fields. Annu. Rev. Neurosci. 31, 479–509.
Huttenlocher, P.R., Dabholkar, A.S., 1997. Regional differences in synaptogenesis in
human cerebral cortex. J. Comp. Neurol. 387, 167–178.
Jann, K., Kottlow, M., Dierks, T., Boesch, C., Koenig, T., 2010. Topographic electrophysiological signatures of FMRI resting state networks. PLoS One 5, e12945.
Ji, D., Wilson, M.A., 2007. Coordinated memory replay in the visual cortex and hippocampus during sleep. Nat. Neurosci. 10, 100–107.
Jouvet-Mounier, D., Astic, L., Lacote, D., 1970. Ontogenesis of the states of sleep in rat,
cat, and guinea pig during the first postnatal month. Dev. Psychobiol. 2, 216–239.
Kerschensteiner, D., Wong, R.O.L., 2008. A precisely timed asynchronous pattern of ON
and OFF retinal ganglion cell activity during propagation of retinal waves. Neuron
58, 851–858.
Khazipov, R., Luhmann, H.J., 2006. Early patterns of electrical activity in the developing
cerebral cortex of humans and rodents. Trends Neurosci. 29, 414–418.
Khazipov, R., Sirota, A., Leinekugel, X., Holmes, G.L., Ben-Ari, Y., Buzsaki, G., 2004. Early
motor activity drives spindle bursts in the developing somatosensory cortex. Nature 432, 758–761.
Kreider, J.C., Blumberg, M.S., 2000. Mesopontine contribution to the expression of active 'twitch' sleep in decerebrate week-old rats. Brain Res. 872, 149–159.
Latsari, M., Dori, I., Antonopoulos, J., Chiotelli, M., Dinopoulos, A., 2002. Noradrenergic innervation of the developing and mature visual and motor cortex of the rat brain: a light and
electron microscopic immunocytochemical analysis. J. Comp. Neurol. 445, 145–158.
Laufs, H., Krakow, K., Sterzer, P., Eger, E., Beyerle, A., Salek-Haddadi, A., Kleinschmidt,
A., 2003. Electroencephalographic signatures of attentional and cognitive default
modes in spontaneous brain activity fluctuations at rest. Proc. Natl. Acad. Sci.
100, 11053–11058.
Leopold, D.A., Logothetis, N.K., 2003. Spatial patterns of spontaneous local field activity
in the monkey visual cortex. Rev. Neurosci. 14, 195–205.
Logothetis, N.K., Wandell, B.A., 2004. Interpreting the BOLD signal. Annu. Rev. Physiol.
66, 735–769.
Lu, H., Zuo, Y., Gu, H., Waltz, J.A., Zhan, W., Scholl, C.A., Rea, W., Yang, Y., Stein, E.A.,
2007. Synchronized delta oscillations correlate with the resting-state functional
MRI signal. Proc. Natl. Acad. Sci. 104, 18265–18269.
Marcano-Reik, A.J., Blumberg, M.S., 2008. The corpus callosum modulates spindle-burst
activity within homotopic regions of somatosensory cortex in newborn rats. Eur. J.
Neurosci. 28, 1457–1466.
Marty, R., Thomas, J., 1963. Electro-cortical responses to stimulation of the cochlear
nerve in the newborn cat. J. Physiol. 55, 165–166.
McCormick, D.A., Trent, F., Ramoa, A.S., 1995. Postnatal development of synchronized
network oscillations in the ferret dorsal lateral geniculate and perigeniculate nuclei. J. Neurosci. 15, 5739–5752.
Mechawar, N., Descarries, L., 2001. The cholinergic innervation develops early and rapidly in the rat cerebral cortex: a quantitative immunocytochemical study. Neuroscience 108, 555–567.
Milh, M., Kaminska, A., Huon, C., Lapillonne, A., Ben-Ari, Y., Khazipov, R., 2007. Rapid
cortical oscillations and early motor activity in premature human neonate. Cereb.
Cortex 17, 1582–1594.
Minlebaev, M., Ben-Ari, Y., Khazipov, R., 2007. Network mechanisms of spindle-burst
oscillations in the neonatal rat barrel cortex in vivo. J. Neurophysiol. 97, 692–700.
Minlebaev, M., Ben-Ari, Y., Khazipov, R., 2009. NMDA receptors pattern early activity in
the developing barrel cortex in vivo. Cereb. Cortex 19, 688–696.
Minlebaev, M., Colonnese, M., Tsintsadze, T., Sirota, A., Khazipov, R., 2011. Early gamma
oscillations synchronize developing thalamus and cortex. Science 334, 226–229.
Mizrahi, E.M., Hrachovy, R.A., Kellaway, P., 2003. Atlas of Neonatal Electroencephalography, Third edition. Lippincott Williams & Wilkins.
Moody, W.J., Bosma, M.M., 2005. Ion channel development, spontaneous activity, and
activity-dependent development in nerve and muscle cells. Physiol. Rev. 85, 883–941.
Muramoto, S., Yamada, H., Sadato, N., Kimura, H., Konishi, Y., Kimura, K., Tanaka, M.,
Kochiyama, T., Yonekura, Y., Ito, H., 2002. Age-dependent change in metabolic response to photic stimulation of the primary visual cortex in infants: functional
magnetic resonance imaging study. J. Comput. Assist. Tomogr. 26, 894–901.
Pampiglione, G., 1977. Development of some rhythmic activity in the E.E.G. of young
pigs, lambs and puppies. Rev. Electroencephalogr. Neurophysiol. Clin. 7, 255–262.
Petersen, J., Diperri, R., Himwich, W.A., 1964. The comparative development of the EEG
in rabbit, cat and dog. Electroencephalogr. Clin. Neurophysiol. 17, 557–563.
U
920
921
922
923
924
925
926
927
928
929
930
931
932
933
934
935
936
937
938
939
940
941
942
943
944
945
946
947
948
949
950
951
952
953
954
955
956
957
958
959
960
961
962
963
964
965
966
967
968
969
970
971
972
973
974
975
976
977
978
979
980
981
982
983
984
985
986
987
988
989
990
991
992
993
994
995
996
M. Colonnese, R. Khazipov / NeuroImage xxx (2012) xxx–xxx
E
10
997
998
999
1000
1001
1002
1003
1004
1005
1006
1007
1008
1009
1010
1011
1012
1013
1014
1015
1016
1017
1018
1019
1020
1021
1022
1023
1024
1025
1026
1027
1028
1029
1030
1031
1032
1033
1034
1035
1036
1037
1038
1039
1040
1041
1042
1043
1044
1045
1046
1047
1048
1049
1050
1051
1052
1053
1054
1055
1056
1057
1058
1059
1060
1061
1062
1063
1064
1065
1066
1067
1068
1069
1070
1071
1072
1073
1074
1075
Please cite this article as: Colonnese, M., Khazipov, R., Spontaneous activity in developing sensory circuits: Implications for resting state fMRI,
NeuroImage (2012), doi:10.1016/j.neuroimage.2012.02.046