ArticlePDF Available

Genetic identification of an embryonic parafacial oscillator coupling to the preBotzinger complex

Authors:

Abstract

The hindbrain transcription factors Phox2b and Egr2 (also known as Krox20) are linked to the development of the autonomic nervous system and rhombomere-related regulation of breathing, respectively. Mutations in these proteins can lead to abnormal breathing behavior as a result of an alteration in an unidentified neuronal system. We characterized a bilateral embryonic parafacial (e-pF) population of rhythmically bursting neurons at embryonic day (E) 14.5 in mice. These cells expressed Phox2b, were derived from Egr2-expressing precursors and their development was dependent on the integrity of the Egr2 gene. Silencing or eliminating the e-pF oscillator, but not the putative inspiratory oscillator (preBötzinger complex, preBötC), led to an abnormally slow rhythm, demonstrating that the e-pF controls the respiratory rhythm. The e-pF oscillator, the only one active at E14.5, entrained and then coupled with the preBötC, which emerged independently at E15.5. These data establish the dual organization of the respiratory rhythm generator at the time of its inception, when it begins to drive fetal breathing.
Genetic identification of an embryonic parafacial
oscillator coupling to the preBo
¨tzinger complex
Muriel Thoby-Brisson1, Mattias Karle
´n2, Ning Wu1, Patrick Charnay3, Jean Champagnat1& Gilles Fortin1
The hindbrain transcription factors Phox2b and Egr2 (also known as Krox20) are linked to the development of the autonomic
nervous system and rhombomere-related regulation of breathing, respectively. Mutations in these proteins can lead to abnormal
breathing behavior as a result of an alteration in an unidentified neuronal system. We characterized a bilateral embryonic
parafacial (e-pF) population of rhythmically bursting neurons at embryonic day (E) 14.5 in mice. These cells expressed Phox2b,
were derived from Egr2-expressing precursors and their development was dependent on the integrity of the Egr2 gene. Silencing
or eliminating the e-pF oscillator, but not the putative inspiratory oscillator (preBo
¨tzinger complex, preBo
¨tC), led to an abnormally
slow rhythm, demonstrating that the e-pF controls the respiratory rhythm. The e-pF oscillator, the only one active at E14.5,
entrained and then coupled with the preBo
¨tC, which emerged independently at E15.5. These data establish the dual organization
of the respiratory rhythm generator at the time of its inception, when it begins to drive fetal breathing.
In mammals, breathing is one of the earliest motor behaviors of the
fetus and it relies on the activity of a brainstem respiratory rhythm
generator (RRG). Recent advances in the neurobiology of breathing in
neonatal mammals suggest that the RRG is located in two prominent
rhythmogenic sites, the preBo
¨tC1and the para-facial respiratory group
(pFRG)2. The genetic regulation of progenitor cell fate and differentia-
tion in the hindbrain, and of the plasticity of the RRG, remains
poorly understood.
The developmental origin and functional nature of the respiratory
rhythm-generating circuits involved in fetal and neonatal breathing
have been studied using mutant mice in which developmental genes
encoding transcription factors have been inactivated, including Egr2,
Phox2b,Hoxa1,Tlx3 and Mafb3–7. In particular, inactivation of Egr2,
the gene for the zinc finger transcription factor Egr2, which controls the
formation of hindbrain rhombomeric segments 3 and 5 (refs. 8,9),
results in defective breathing. This is attributed to a reorganization of
neuronal circuits in the caudal pontine reticular formation that
consequently leads to poor survival at birth6. Further evidence from
chicks shows that Egr2 is involved in the specification of a central
rhythm generator at the level of the facial motor nucleus10. Although
these studies have shown that the importance of hindbrain segmenta-
tion extends beyond modular anatomical organization to the
level of network assembly and function, they have failed to identify a
candidate Egr2-derived cell group that would explain the mutant
respiratory deficit.
Phox2b is a transcription factor that is specifically expressed and
required in neurons that form the visceral reflex circuits controlling
digestive, cardiovascular and respiratory functions11,12. In humans, a
heterozygous mutation in PHOX2B is the main cause of congenital
central hypoventilation syndrome (CCHS)13,14, a genetic disease that
typically manifests itself at birth by respiratory distress during sleep15.
Newborn mice that are heterozygous for the most common human
mutation have a slowed and often irregular breathing pattern, do not
respond to hypercapnia, and die at birth from respiratory failure16.
Anatomically, these mutants selectively lack a group of glutamatergic,
Phox2b-expressing interneurons in the ventro-lateral medulla in the
vicinity of the facial motor nucleus16 called the retrotrapezoid nucleus
(RTN)17. The RTN has previously been identified as a CO2/pH sensor
system in the adult rat18,19. Notably, the rhythmogenic pFRG over-
laps anatomically with the RTN and was recently shown to host
CO2-sensitive Phox2b-positive interneurones in rat neonates20, thus
supporting the view that the pFRG and RTN may correspond to
neonatal and adult forms of the same neuronal population19.
We investigated embryonic stages using knock-in alleles of Egr2 and
identified a population of Phox2b-positive interneurons deriving from
Egr2-expressing cells that forms an e-pF oscillator. The e-pF oscillator,
which is first active at E14.5, couples with the developing preBo
¨tC
oscillator within 24 h, thus establishing the dual organization of the
RRG at the time at which it begins to pace fetal breathing.
RESULTS
The embryonic parafacial oscillator
We used mice carrying a knock-in of the Cre recombinase gene into the
Egr2 locus (Egr2cre)21 and a Cre-responsive indicator allele (R26R-
EYFP)22 to trace the derivatives of Egr2-expressing cells. At E14.5, cells
derived from Egr2-positive progenitors formed two transverse stripes
with little cell dispersal along the anterior-posterior axis; these stripes
correspond to the location of and are likely to be derived from the
Received 14 April; accepted 29 May; published online 5 July 2009; doi:10.1038/nn.2354
1Institut de Neurobiologie Alfred Fessard, Centre National de la Recherche Scientifique UPR2216, Gif sur Yvette, France. 2Department of Cell and Molecular Biology,
Karolinska Institute, Stockholm, Sweden. 3Institut National de la Sante
´et de la Recherche Me
´dicale, U784, Ecole Normale Supe
´rieure, Paris, France. Correspondence
should be addressed to G.F. (gilles.fortin@inaf.cnrs-gif.fr).
1028 VOLUME 12
[
NUMBER 8
[
AUGUST 2009 NATURE NEUROSCIENCE
ARTICLES
© 2009 Nature America, Inc. All rights reserved.
rhombomeric 3 and 5 segmental domains that have been observed
during hindbrain development (see ref. 9, the r5 stripe is shown in
Fig. 1a). Notably, inspecting the ventral surface of the hindbrain, this
restriction along the anterior-posterior axis was clearly disrupted
laterally, resulting in a parafacial stripe of yellow fluorescent protein
(YFP)-expressing cells caudal to r5, flanking and partially capping the
lateral aspect of facial motor nucleus (nVII) and continuing approxi-
mately 200 mm caudal to the nVII (Fig. 1a). Using optical recordings
from brainstem en bloc preparations23 after Calcium Green 1AM
loading, we identified a rhythmic cell population that matched this
parafacial YFP+territory (frequency, 14.3 ± 3.9 bursts min1;range,
8–18 bursts min1;n¼19 preparations; Fig. 1b–f). Occasionally, a
burst of activity in the neighboring nVII (Fig. 1d) was associated with a
burst of activity in the parafacial domain (see below). Low-resolution
imaging revealed bilateral coactivation of parafacial domains and the
absence of activity in other regions of the hindbrain. In transverse facial
slices (n¼8), a comparable rhythm was maintained in cells of the
parafacial region (Supplementar y Fig. 1). These data demonstrate that
the parafacial domain forms an intrinsically active rhythmic network,
which we refer to as the e-pF oscillator.
Imaging at cellular resolution showed synchronized activity of
individual e-pF cells (Fig. 1e,f) and synchronized rhythmic changes
in fluorescence were detected in 95% of the YFP+cells (98 of 103 cells
from three en bloc preparations) that were sampled along the rostro-
caudal extent (B700 mm) of the e-pF (Fig. 1g,h). Because the
proportion of YFPcalcium-loaded cells in the e-pF area was less
than 10%, this suggests that almost all of the e-pF cells expressed YFP
and were therefore presumably derived from rhombomeric segments 3
or 5. Active cells were not detected before E14.5 (data not shown; E12.5,
n¼3; E13.5, n¼5); thus, E14.5 marks the onset of the e-pF.
On the basis of previous work showing that respiratory-related
Phox2b-positive and neurokinin-1 receptor (NK1R)–positive neurons
are located in an area similar to the e-pF at E15.5 (ref. 16), we examined
those markers in the e-pF. We selected rhythmically active e-pF neurons
by loose patch recording, filled them with biocytin and determined
whether the Phox2b and NK1R proteins were present by immuno-
labeling. We found that 15 of 16 e-pF cells expressed Phox2b (Fig. 1i)
and 7 of 10 expressed NK1R (Fig. 1j); thus, a large fraction of e-pF cells
expressed both markers at E14.5. The NK1Rs were functional because
the e-pF rhythm frequency increased twofold in the presence of substance
P, the endogenous ligand for NK1R (0.1 mM, data not shown). In
Egr2cre/+;R26R-EYFPembryos, most, if not all, of the YFP+cells in the
parafacial area expressed Phox2b and could be easily distinguished
from the Phox2b+,YFP
and Islet1/2+nVII cells (Fig. 1k). All Phox2b+
cells in the parafacial region express the type 2 vesicular glutamate
transporter (VGlut2) at birth16, indicating the probable glutamatergic
nature of e-pF cells. The number of e-pF YFP+
,Phox2b
+, Islet1/2
neurons on one side (280 ± 36, n¼3) was similar to the number of
rhythmic cells detected in calcium-loaded preparations (259 ± 54 cells,
n¼5), indicating that the e-pF oscillator is composed of Phox2b+
neurons, which are probably glutamatergic, derived from Egr2-expres-
sing precursors.
Operating principles of the e-pF oscillator
We then investigated the cellular properties underlying rhythm gen-
eration in the e-pF oscillator. In whole-cell recordings performed on
E14.5 whole hindbrain preparations, rhythmic burst discharges of
action potentials in e-pF cells (43 of 43; Fig. 2a) appeared as all-
or-none voltage-dependent events. Spontaneous bursts were curtailed
by short negative-current pulses (Fig. 2b) that were applied during the
burst, whereas burst discharges were evoked in between spontaneous
bursts by comparable current pulses of opposite polarity (Fig. 2c).
Slowly ramping the somatic potential of e-pF cells from 80 to
+40 mV revealed a tetrodotoxin-sensitive (n¼3, data not shown)
and riluzole-sensitive (n¼10) persistent sodium current (INaP;Fig. 2d).
Inthepresenceofriluzole(20mM),rhythmicactivityofthee-pFcell
F/F
F/F
10%
F/F
4%
5 s
5 s
F/F50
10
1
0
12
10
1
e-pF
A
r5
YFP IsI1/2
YFP IsI1/2 Phox2b
12
3
4
5
67
9
8
11 10
12
YFP Ca Green1 Merge
Phox2b Biocytin Merge
NK1R Biocytin Merge
1
10
Direct F
M
D
M
abc
e
g
i
j
k
h
f
d
Figure 1 A parafacial oscillator emerges at E14.5 and is derived from Egr2-
positive territories in the mouse embryo hindbrain. (a) Partial ventral view of
a hindbrain from an Egr2cre/+; R26R-EYFP embryo showing the respective
positions of cells derived from rhombomeric segments 3 and 5 (YFP positive,
green) and nVII motoneuronal populations (Islet1/2 positive, red). Note the
stripe of YFP-expressing cells caudal to rhombomeric segment 5, flanking the
lateral aspect of the nVII (blue outline). (bd) A Calcium Green 1AM–loaded
whole hindbrain preparation (b) showing fluorescence changes, which were
generally restricted to the e-pF oscillator (red outline) in c,andwere
sometimes concomitant to activity of the nVII (d). (e,f) Photomicrograph of
e-pF cells (numbered 1 to 12) during a burst of activity (e) and corresponding
individual (black) and average (red) relative fluorescence changes traces (f).
(g) Same field of an Egr2cre/+; R26R-EYFP preparation showing YFP-
expressing cells (red), Calcium Green 1AM–loaded cells (green) and the
merged image used to derive the individual rhythmic activities of ten double-
labeled (yellow) cells. (h) Fluorescence changes of individual cells (black)
and averaged signal (red) from g.(i,j) Immunolabeling for Phox2b (i) and for
NK1R protein (j) in two biocytin-filled e-pF neurons. (k) Single transverse
section from an E14.5 Egr2cre/+; R26R-EYFP embryo that were triple
immunolabeled with antibodies specific to Islet1/2 (blue), Phox2b (red) and
YFP (green). Cells of the e-pF expressing both Phox2b and YFP are shown in
yellow and nVII motoneurons expressing both Phox2b and Islet1/2 appear in
purple. Scale bars represent 20 mm(e,g,i,j) and 200 mm(ad,k). A, anterior;
D, dorsal; M, median.
NATURE NEUROSCIENCE VOLUME 12
[
NUMBER 8
[
AUGUST 2009 1029
ARTICLES
© 2009 Nature America, Inc. All rights reserved.
population was abolished and the all-or-none bursts of action poten-
tials that were evoked by depolarizing current pulses or that occurred
spontaneously were replaced by single spikes (Fig. 2e). Hyperpolariza-
tion-activated and cyclic nucleotide–gated cation channel (HCN)-
mediated currents (Ih), which are often expressed together with INaP
in pacemaking neurons24, were present in 25 of 26 e-pF neurons
(Fig. 2f). Although the hyperpolarization level required to elicit a
somatic Ihcurrent was typically 20 mV below the resting membrane
potential of e-pF cells (Vm¼47 ± 1.5 mV, n¼20), application of
ZD7288 (100 mM) systematically blocked the Ihcurrent present in e-pF
cells and reduced the frequency of the e-pF oscillator from 13.7 ± 1.3
bursts min1to 5.4 ± 0.7 bursts min1(n¼11; Fig. 2g), suggesting a
role for possibly distally located HCN channels in the modulation
of the rhythm.
Rhythm generation and intercellular synchrony of the e-pF oscillator
were preserved in the presence of 10 mMCNQX,whichblocksgluta-
matergic transmission mediated by AMPA/kainate receptors (n¼5;
Fig. 2h), as well as in the presence of either the m-opioid agonist
D-Ala2-N-Me-Phe4-glycol5-enkephalin (DAMGO, 0.3 mM, n¼5) or a
cocktail of bicuculline (10 mM) and strychnine (5 mM), which block
GABAAand glycinergic receptors, respectively (n¼5, data not shown).
Furthermore, the e-pF oscillator rhythm was preserved (although
slower) in mice that do not express VGlut2 (Vglut2f/f;PCre; exons 4–6
of Vglut2 are flanked with loxP sites and cre is driven by the Pgk
promoter)25 (n¼3; Fig. 2i). Moreover, the
e-pF oscillator activity was spared in the pre-
sence of the NMDA-R antagonist D-2-amino-
5-phosphonovaleric acid (AP5, 5 mM, n¼6; Fig. 2j). These data
suggest that glutamatergic synaptic transmission is not essential for
rhythm generation or for intercellular synchrony of the e-pF oscillator.
Application of lanthanum (La3+,100mM, n¼4), which can efficiently
block hemichannels, but not gap junctions permeabilities26,27, failed to
alter the e-pF collective synchronous activity (Fig. 2k). However,
carbenoxolone (CBX, 50 mM, n¼11) blocking gap junctions, in
addition to hemichannels, abolished intercellular synchrony (Fig. 2l).
Under CBX, e-pF cells that maintained rhythmic fluorescence changes
(96 of 229 e-pF cells from ten preparations) were readily silenced by
further application of riluzole (n¼5; Fig. 2l). We observed similar
effects of CBX (n¼2), riluzole (n¼3), CBX and riluzole
(n¼2), and ZD7288 (n¼3) on the activity of e-pF cells in transverse
facial slices (five slices, data not shown), further establishing the e-pF
as the source of rhythmic activities.
Low-resolution imaging on E14.5 whole hindbrain preparations
indicated that application of CNQX (n¼3) spared rhythmic activities
and intercellular synchrony in parafacial domains on either side of
the midline, but caused their left/right de-synchronization. This de-
synchronization was not observed in response to the bicuculline/
strychnine cocktail (n¼3). Independent rhythms in left and right
parafacial domains were also observed in transverse slices from the axial
level of the nVII (n¼4; Supplementary Fig. 2). Thus, on the one hand,
e-pF cells collectively function as an oscillator in a manner that is
Neuron
e-pF int
lh current lh current l/V curve
Vm (mV)
e-pF int
e-pF
12
1
1
1
1
1
e-pF
12
e-pF
e-pF
e-pF
e-pF
11
9
10
1
1
10
9
–60 +20–40 – 20 0 –60 +20–40 –20
–120 – 100
CTL (n = 14)
ZD (n = 5)
–80 –60
10
–10
0
–20
–30
–40
0
Ril 50 pA
Voltage (mV)
0
20
40
CTL–Ril
Current (pA)
lh amplitude (pA)
Control
–60 mV
2 s
i
V
–50 mV
i
i
20 mV
1 s
1 s
–100 V
–50 mV
50 pA
–0.1 nA
–60 mV
F/F
10%
F/F
10%
F/F
2%
F/F
4%
F/F
4%
F/F
4%
F/F
4%
–45 mV
Riluzole
CTL
V
V
CTL ZD
13 mV s–1
Neuron
i
Neuron
Control
ZD7288
WT in CNQX
WT
Control
Control
AP5
VGlut2f/f;PCre
CBX + riluzole
CBX
La3+
Control Control
e-pF int
20 mV
20 mV
–50 mV
0.4 nA
20 mV
–50 mV
5 s
20 mV
0.1 nA
20 mV
0.1 nA
1 s
5 s
5 s
5 s
5 s
5 s
5 s
2 s
V
i
a
d
f
hi
jk
l
g
e
bc Figure 2 Operating principles of the e-pF
oscillator at E14.5. (a) Spontaneous rhythmic
depolarization and firing (top trace) of an e-pF cell
synchronous with e-pF population integrated
activity (bottom trace). (b,c) Spontaneous bursts
(top, gray control trace) in phase with the
population activity (bottom traces, e-pF int) could
be prematurely terminated (b, black traces) or
fully evoked (c, black traces) by negative and
positive current pulse injections (i, middle traces),
respectively. (d,e) Bursting activity in e-pF cells
relied on the INaP current. Riluzole (Ril) blocked
the control (CTL) inward current activated in e-pF
cells by slow depolarizing voltage ramps (d, left).
The INaP current-voltage relationship was obtained
by subtracting the riluzole-resistant current from
the control current (d, right, CTL – Ril). Riluzole
also blocked spontaneous and evoked burst of
activity in the e-pF (e). (f,g) The hyperpolarization-
activated current Ihpresent in e-pF cells was
blocked by ZD7288 (ZD, f) and modulated
the e-pF rhythm (g). Error bars represent s.e.m.
(hj) Glutamatergic transmission was not required
for rhythmic activity of the e-pF oscillator. The
e-pF rhythm was maintained in 10 mMCNQX(h),
in Vglut2f/f;PCre mutant mice25 preparations (i)and
in the presence of 5 mM AP5 (j). (k,l) Intercellular
coactivation in the e-pF oscillator relied on gap
junctions. Intercellular coactivation was maintained
after blockade of hemichannel permeabilities by
100 mM lanthanum (La3+, control not shown, k),
but was lost after CBX (50 mM) treatment and
active cells were silenced on further application
of riluzole (l). Black traces in glrepresent
fluorescence changes in individual e-pF cells
captured in the frame and gray traces represent
the average fluorescence changes over the regions
encompassing these cells.
1030 VOLUME 12
[
NUMBER 8
[
AUGUST 2009 NATURE NEUROSCIENCE
ARTICLES
© 2009 Nature America, Inc. All rights reserved.
largely independent of glutamatergic transmission; on the other hand,
e-pF bilateral coactivation appears to rely on glutamatergic commis-
sural projections that are already established at E14.5.
The e-pF oscillator couples with the preBo
¨tC oscillator
In mice, the respiratory preBo
¨tC oscillator, located at the vagal level of
the hindbrain, is first active at E15.5 and can maintain the rhythmic
activity of hypoglossal motoneurons in transverse slices28. Moreover, at
E15.5, fluorescence changes in the e-pF area and the nVII occurred
together with bouts of hypoglossal nerve (XIIn) activity in a conti-
nuous rhythmic and synchronized manner, indicating that there is a
common source of rhythmic activity, which possibly involves the e-pF
oscillator (Fig. 3ac). To investigate a possible interaction between the
e-pF and the preBo
¨tC, we carried out simultaneous optical recordings
of the e-pF oscillator and nVII and electrophysiological recordings of
XIIn on E15.5 preparations.
We first investigated the consequences on nVII and XIIn activities of
a transection that was just posterior to the facial motor nucleus, in
between the e-pF oscillator and the preBo
¨tC oscillators (n¼4).
Notably, rhythmic activities were maintained on both sides of the
section. Caudal to the section, the XIIn had a lower frequency than
intact preparations (Fig. 3a,d), reflecting the presence of an active
preBo
¨tC oscillator. Indeed, this activity was abolished by bath applica-
tion of DAMGO29 (n¼3, data not shown), but was unaffected by
riluzole (n¼3), which also failed to modify the rhythm observed in
E15.5 preBo
¨tC transverse slices (n¼6; Fig. 3d). Rostral to the section,
the e-pF oscillated faster than in intact preparations, whereas the
adjacent nVII was completely silenced, indicating that the e-pF prob-
ably does not entrain it directly (Fig. 3a). The increased frequency of
the e-pF was consistent with that observed in facial transverse slices,
suggesting that partial e-pF cellular loss and/or severed commissural
fibers are causal factors. At any rate, caudal to the section, the resulting
slower rhythm suggested the elimination of a descending excitatory
input, indicating that the two oscillators interact at E15.5.
Second, in intact E15.5 preparations, bath applications of riluzole
that selectively suppress the e-pF activity (n¼8) led to maintained
rhythmic activities in the nVII and XIIn, although at about half of the
frequency observed in control conditions (Fig. 3b,d). These data
indicate that functional impairment of the e-pF results in a slower
rhythm that is driven by the sole preBo
¨tC oscillator. Finally, silencing
the preBo
¨tC oscillator by application of CNQX (n¼5, Fig. 3c,d)or
DAMGO (n¼7, Fig. 3d) led to a selective cessation of rhythmic
activities of the nVII and XIIn, demonstrating the pre-motor status of
the preBo
¨tC oscillator. Our data support the view that the e-pF
oscillator increases the frequency of rhythmic motor bursts generated
by the preBo
¨tC. Altogether, these data suggest that the dual organiza-
tion of the RRG described at later stages2,30 is already achieved by the
time of its inception at E15.5.
We then investigated, during the E14.5–15.5 period, the establish-
ment of the continuous rhythmic motor output. We confirmed that
spontaneous collective cellular rhythmic behavior was absent at E14.5
in transverse preBo
¨tC slices (n¼5), denoting its immature status28.In
some of the E15.5 and in all of the E14.5 whole hindbrain preparations,
we observed discrete failings of the nVII and XIIn bursts, despite the
continuous generation of e-pF rhythmic bursts (Fig. 4a). This was
reminiscent of skipped respiratory cycles, as opioid-induced reduction
of excitability in one (preBo
¨tC) of the two coupled oscillators causes
transmission failures of the rhythmic drive from the other coupled
oscillator in the neonatal RRG31. Therefore, we investigated, during
E14.5–15.5, the possibility that excitability build-up in the preBo
¨tC in
the presence of an already active e-pF rhythmic drive may conversely
lead to discrete occurrences of motor bursts in phase with e-pF bursts.
To explore this, we first compared the frequency distribution of motor
bursts with that of e-pF bursts between E14.5 and E15.5 prepara-
tions. At E14.5, the frequencies of motor bursts (3.4 ± 1.5 burstsmin1,
n¼11) and e-pF bursts (10.6 ± 1.7 bursts min1,n¼11) were
significantly different (Po0.001), whereas the distributions were
found to partially overlap at E15.5 (Fig. 4b). This resulted from an
increased occurrence of motor bursts (8.6 ± 2.1 bursts min1,n¼9),
whereas the frequency of the e-pF oscillator was unchanged (10.2 ±
2.4 bursts min1,n¼9). At E14.5, each individual burst of motor
activity was generated with a delay after the onset of an e-pF burst,
which was reduced or non-existent by E15.5 (Fig. 4c,d). These data
indicate that the continuous generation of rhythmic motor bursts
Figure 3 Coupled oscillators control the motor
activity at E15.5. Simultaneous optical
recordings of the e-pF oscillator and nVII
activities and electrophysiological recordings of
XIIn were performed at E15.5 in en bloc
preparations. Note that all activities are
synchronized in control conditions (top set of
traces in ac). (a) A transverse section below the
nVII led to spared independent rhythmic
activities of the e-pF oscillator (top trace) and
the XIIn (bottom trace) and to complete
suppression of activity in the nVII (middle
trace), suggesting that the preBo
¨tC oscillator
drove motoneuronal pools before the section.
(b) Selective silencing of the e-pF oscillator
by 20 mM riluzole resulted in a slower motor
rhythm. (c) CNQX application (preserving
left/right de-synchronized activities of the e-pF)
disrupted the activity of the preBo
¨tC oscillator
and abolished motor activities. (d)Graph
representing the percentage change (mean ±
s.e.m.) of the frequency of rhythmic activities of
the e-pF (black bars), nVII or XIIn (Motor, gray
bars) and preBo
¨tC (white bar) after the transverse section (n¼4), the applications of 10 mMCNQX(n¼11), 0.3 mM DAMGO (n¼7) and riluzole (Ril,
n¼8) in whole-hindbrain preparations (WHB) or transverse slice (slice) preparations (n¼6). * Po0.05.
Control
Control
CNQX
Section 2% 2%
2%
10 s 10 s
10 s
e-pF
e-pF
e-pF
nVII
nVII
nVII
Xlln
Control
Riluzole
e-pF
nVII
Xlln
e-pF
e-pF
Motor
PreBötC
nVII
Xlln
200
Percentage of control
150
Section CNQX DAMGO Ril Ril DAMGO
SliceWHB
Xlln
*
*
***
*
*
100
50
0
Xlln
e-pF
nVII
Xlln
Xlln
ab
dc
NATURE NEUROSCIENCE VOLUME 12
[
NUMBER 8
[
AUGUST 2009 1031
ARTICLES
© 2009 Nature America, Inc. All rights reserved.
does not appear abruptly at E15.5 together with the preBo
¨tC oscillator,
but instead results from a dynamic process in which the e-pF oscillator
seems to be important. In fact, the frequency of motor activities
increased linearly with the value of transmission ratios; that is, the
number of motor bursts to the number of e-pF bursts (E14.5, n¼11;
E15.5, n¼19) to reach unity at E15.5 in 12 of 19 preparations (two
cases shown in Fig. 4e). Inspection of activities in individual prepara-
tions clearly illustrated the quantal nature of motor burst incorpora-
tion, eventually giving way to a continuous rhythm (Fig. 4c). This
dynamic process reflected an increasing excitability at pre-motor/
motor synapses during E14.5, as the time lags separating the onsets
of e-pF and motor bursts decreased with increasing transmission ratios
and eventually approached zero when the rhythm became continuous
(12 of 19 preparations at E15.5 and 5 of 5 preparations at E16.5;
Fig. 4d,f). Therefore, the motor rhythm became continuous during
the E14.5–15.5 period, as a result of increased efficiency of the e-pF in
transmitting activity to motoneuronal pools via a maturing preBo
¨tC,
which acquires autonomous rhythm generation at the end of the process.
We next investigated the signature of this coupling among oscillators
at the single-cell level by recording the membrane potential of indivi-
dual e-pF cells and the activity of the preBo
¨tC oscillator in E15.5 en
bloc preparations (Fig. 5). We found, using low chloride concentration
(6 mM) pipette solution for whole-cell recordings, that the e-pF oscillator
comprised two distinct sets of neurons. During a burst of activity in
the preBo
¨tC, 16 of 30 e-pF cells discharged a burst of action potentials
(Fig. 5a). In contrast, the remaining 14 e-pF cells showed a pause in
firing that was associated with a barrage of chloride-mediated synaptic
potentials (Fig. 5b). Because these differences could be the results of a
variable effectiveness in changing the intracellular chloride concentra-
tion through whole-cell recordings, we performed voltage-clamp
experiments to check the polarity and kinetics of underlying synaptic
currents. Working at a holding potential of 40 mV, excited (n¼7)
and pausing (n¼10) e-pF cells featured both inward and outward
background synaptic events (Fig. 5b,e), and during preBo
¨tC bursting
activity, they featured barrages of prominently fast (decay, B1.5 ms)
inward (Fig. 5b,c) and slow (decay, B8.0 ms) outward (Fig. 5e,f)
synaptic currents, respectively. Fast inward synaptic currents were
blocked by CNQX (data not shown) and slow outward synaptic
currents by bicuculline (Fig. 5g), indicating their respective mediation
by AMPA/kainate and GABAAreceptors. Thus, pausing cells may
eventually show active inhibition during preBo
¨tC inspiratory
bursts, whereas those discharging in phase with the preBo
¨tC are
candidate neurons through which the e-pF oscillator could contribute
to increase the frequency of the RRG. These data suggest that synaptic
interactions are established at E15.5 between e-pF and preBo
¨tC
oscillators. In sum, these experiments show that the e-pF oscillator is
fated to couple with the preBo
¨tC, the essential oscillatorinvolved in the
control of respiration.
Elimination of the e-pF oscillator in Egr2 null mutants
Our data indicate that the e-pF may be important in increasing
the frequency of fetal breathing. In Egr2lacZ/lacZ mice, which are null
mutants for Egr2, the elimination of populations derived from
rhombomeric segments 3 and 5 leads to abnormally slow breathing
at birth and poor survival6. Calcium imaging in E15.5 Egr2lacZ/lacZ
preparations (Fig. 6) showed rhythmic fluorescence changes of
the nVII in phase with the XIIn activity, although at half of the
frequency of Egr2lacZ/+ or wild-type embryos (pooled Egr2lacZ/+/WT,
f¼8.9 ± 0.7 bursts min1,n¼7; Fig. 6ac;Egr2lacZ/lacZ,f¼3.8 ± 0.3
bursts min1,n¼9; Fig. 6hj).
Notably, there was a complete absence of rhythmic fluorescence
changes in the e-pF area in the homozygous mutants (Fig. 6i,j).
Anatomically, the longitudinal stripe of NK1Rexpression lateral to the
nVII (Fig. 6d,e) was absent (Fig. 6k,l). In transverse and parasagittal
(data not shown) sections, NK1R expression (Fig. 6f) and Phox2b+/
Islet1/2cells (Fig. 6g) were lacking at the location of the e-pF
e-pF
e-pF
2%
10 s
10 s 1 s
F/F
nVII
nVII
e-pF
nVII
e-pF
15
1
0
3
3
1
1
22
10
0 0.2 0.4 0.6
Transmission ratio
0.8 1.0 0 0.2 0.4 0.6
Transmission ratio
0.8 1.0
5
0
nVII
Xlln
E14.5
E14.5
1%
E14.5
E15.5
E15.5
Freq. (min–1)
Mot. Mot.e-pF e-pF
100
0 5 10 15 0 5 10 15
Nb of preparation (%)
Mot. freq. (min–1)
Lag e-pF-nVII (s)
50
0
a
b
ef
cd
Figure 4 Coupling between the e-pF oscillator and motor activity during
development. (a) Spontaneous calcium changes in the e-pF (red trace) and
nVII (blue trace) with electrophysiological recording of XIIn (black trace)
in an E14.5 preparation. Note the absence of motor bursts with continuous
rhythmic e-pF bursts. (b) Distributions of the frequencies of motor (empty
bars, blue fit line) and e-pF (black bars, red fit line) activities obtained from
E14.5 (left graph, n¼11) and E15.5 (right graph, n¼9) preparations. Grey
bars indicate overlapping bins. (c,d) Calcium changes in the e-pF (red traces
and triangles) and nVII (blue traces and triangles) in two (top and middle)
E14.5 and one (bottom) E15.5 preparation (c). Note the increased
occurrence of motor bursts coupled to the e-pF bursts, associated with the
reduction of the time lag (distance between downward red and blue arrows
and vertical lines) between the sequential onsets of the e-pF and nVII bursts (d).
(e,f) Developmental trends among E14.5 (filled circles) and E15.5 (empty
circles) preparations; motor frequency increased (e) and the e-pF–nVII time
lag decreased (f) with augmenting transmission ratios. Data points 1, 2
and 3 correspond to the top, middle and bottom traces in c.
1032 VOLUME 12
[
NUMBER 8
[
AUGUST 2009 NATURE NEUROSCIENCE
ARTICLES
© 2009 Nature America, Inc. All rights reserved.
oscillator (Fig. 6g,n). Moreover, sectioning the preparation caudal to the
nVII reduced the frequency of rhythmic activity by half in Egr2lacZ/+
preparations, as shown above for wild-type embryos (f¼5.1 ± 0.2 bursts
min1,n¼16 pooled genotypes), but had no significant effect on
the rhythm in Egr2lacZ/lacZ mutants (f¼4.5 ± 0.4 bursts min1,n¼5,
P¼0.3). In fact, the slow rhythm of the homozygous mutants was
comparable with that produced in preBo
¨tC transverse slices prepared
from embryos of either genotypes (wild type, 5.1 ± 0.8 bursts min1,
n¼6; Egr2lacZ/lacZ, 5.6 ± 1.5 bursts min1,n¼5; P¼0.9). Egr2 is thus
required for the development of the NK1R- and Phox2b-expressing
cells that form the e-pF oscillator. This result suggests that the reduced
breathing frequency observed in newborn Egr2 homozygous mutant
pups6may be a result of a lack of entrainment of the preBo
¨tC by
the e-pF.
Figure 6 The e-pF oscillator is absent in Egr2 null
mutant embryos. (a,b) Ventral view of an E15.5
wild-type (WT) hindbrain (a) over the facial area
(inset) imaged at higher magnification (b) during a
burst of activity of the e-pF (red outline) and the
nVII (blue outline). (c) Traces showing the fast
and synchronized rhythmic fluorescence changes
of the e-pF (red) and nVII (blue) with electrical
activity of the XIIn (black). (d,e) Whole-mount
double immunolabeling for NK1R (green) and
Islet1/2 (red) over one facial area (d,NK1Ronly).
The inset (white rectangle) shows the lateral
aspect of the nVII at a higher magnification (e).
(f,g) Single transverse sections, taken at the level
indicated by the arrow in d, were double immuno-
labeled for NK1R (green) and Islet1/2 (red) in f,
and for Phox2b (green) and islet1/2 (red) in g.
(hn)Egr2 null mutant (data are presented as
in ag) showing the more rostral position of the
facial area, owing to the absence of rhombomeric
segment 5 (h), and the lacking activity of the
e-pF (i,j) associated with maintained, but slower,
synchronous rhythmic motor activities of the
nVII and XIIn (j). The absence of the e-pF in the
mutant was associated with a loss (arrowheads) of NK1R+and Phox2b+neurons lateral and ventral to the nVII (k,l,n). Dotted lines in aand hindicate the
caudal limit of migrating pontine neurons. Scale bars represent 200 mm.
e-pF cell
20 mV
0.5 s
20 mV
0.5 s 0.2 s
0.2 s
50 pA
50 pA
40 pA
Control
Bic
10 ms
40 pA
10 ms
= 1.72 ± 0.06 ms
= 8.85 ± 0.76 ms
(ms)
= 1.78 ± 0.23 ms
= 7.39 ± 0.23 ms
Vh = –40 mV
Vh = –40 mV
e-pF cell
–50 mV
–50 mV
PreBötC
int
PreBötC
int
25
20
15
Frequency (%)
10
5
0
25
20
15
Frequency (%)
10
5
0
0 2 4 6 8 10 12 14 16
(ms)
0 2 4 6 8 10 12 14 16
abcg
def
Figure 5 The e-pF oscillator at E15.5 comprises two types of neurons. (a,d) Membrane potential trajectories (top traces) of two representative e-pF cells at
E15.5 and preBo
¨tC population activity (bottom traces) reveal that e-pF neurons were either excited (a), discharged a burst of action potential or were inhibited
(d), showing a pause in firing, during preBo
¨tC rhythmic bursts. Red traces show the average membrane potential trajectory and integrated activity of the
preBo
¨tC corresponding to ten superimposed cycles (black traces) locked on the onset of the preBo
¨tC bursts (vertical gray lines and arrowheads). (b,e) In voltage
clamp mode, during preBo
¨tC bursts of activity (gray background), excited (b) and pausing (e) e-pF cells showed a barrage of prominently inward (black
triangles) and outward (white triangles) synaptic currents, respectively. (c) In excited e-pF cells, inward synaptic currents had fast kinetics. The top set of
traces shows five superimposed inward currents (black triangle, black traces) and their average (red trace), which we used to derive the decay time constant (t)
after single exponential fit. The bottom set of traces shows the same analysis for slow outward events collected in between bursts (white triangle). (f)Datafora
pausing e-pF cell are presented as in c.(g) Histograms showing that the bimodal frequency distribution of t’s in one pausing e-pF cell (n¼118 events) in
control (top histogram) transformed under bicuculline into a modal distribution (bottom histogram) owing to preservation of fast (n¼132 events), but not
slow, synaptic events. Whole-cell recordings were performed using a low (6 mM) chloride pipette solution.
F/F
A
M
D
M
F/F
F/F 2%
F/F 4%
10 s
10 s
0
4
0
2
Wild type
Egr2
–/–
e-pF
Xlln
nVll
e-pF
Xlln
nVll
cg
h
j
kl m
n
i
abdef
NATURE NEUROSCIENCE VOLUME 12
[
NUMBER 8
[
AUGUST 2009 1033
ARTICLES
© 2009 Nature America, Inc. All rights reserved.
DISCUSSION
We have identified a neuronal population in the mouse hindbrain that
shows spontaneous rhythmic activity as early as E14.5. The e-pF
oscillator, composed of a small number of neurons (B300 per side),
constitutes, to the best of our knowledge, the earliest hindbrain
neuronal population showing a continuous spontaneous rhythmic
activity (with a period in the second range) that affects the develop-
ment of the RRG. Cells of the e-pFoscillator are confined to the surface
of the hindbrain and are derived from Egr2-expressing cells. We also
found that the e-pF oscillator cells expressed the homeodomain factor
Phox2b, a transcription factor that is expressed and required in
neuronal types that maintain bodily homeostasis through the reflex
control of digestive, cardiovascular and respiratory functions32.Inthe
rat, Phox2b is expressed by the chemosensitive cells of the pFRG in
neonates20 and of the RTN in adults18. Our data now extend the role of
Phox2b in the specification of visceral reflex circuit11,fromfirstand
second order sensory neurons (including the RTN19), noradrenergic
centers or efferent elements, to a set of interneurons linked to the
inception of the fetal respiratory rhythm.
Recurrent glutamatergic synaptic inputs are essential for establish-
ment of the rhythmic inspiratory drive in preBo
¨tC neurons1,25,33,34.
This is not the case in the e-pF, where we found that rhythm generation
was maintained in several conditions that impaired glutamatergic
transmission. We provide pharmacological evidence that the e-pF
rhythm at E14.5 relies on INaP and is modulated by Ih, although the
cellular distribution of HCN channels and the allosteric regulations35
ensuring their activation in e-pF cells remain unknown. Furthermore,
intercellular synchronization in the parafacial region appears to require
communication through gap junctions (CBX sensitive) and probably
not through hemichannels (La3+ insensitive). This contrasts with
synchronization of parafacial regions across the midline that relies on
glutamatergic synaptic transmission mediated by AMPA/kainate recep-
tors. Because bilateral coactivation is absent in facial transverse slices,
the commissural system involved probably resides beyond the slice
limits. One may argue that the preBo
¨tC, which is sensitive to CNQX1,
has built-in commissural connectivity36,37, and is entrained by and
couples to the e-pF, could take part in setting up this synchrony.
However, no functional commissural connectivity was found for the
preBo
¨tC at E14.5, when bilateral coactivation of the e-pF is overt28.
Hence, the possibility remains that there is another unknown com-
missural system or, more simply, that some of the glutamatergic e-pF
cells bear commissural axons coursing outside the anterior-posterior
limits of the slice.
The preBo
¨tC oscillator is spared in Egr2 null mutants, in which the
e-pF oscillator does not form, indicating that the latter is not required
for the emergence of the former. Thus, together with previous evidence
that the preBo
¨tC can develop in an isolated context38, our data indicate
that rhythmogenic circuits in the vicinity of branchiomotor nuclei
(e-pF/nVII, preBo
¨tC/nucleus ambiguous) are deployed independently
at pre- and post-otic levels and connect to form the RRG. The indepen-
dent emergence of the e-pF and preBo
¨tC is consistent with speculations
that they have distinct evolutionary origins; the pFRG would have
appeared first during the evolution of vertebrates, possibly co-opted
when abdominal expiration was combined with buccal pumping in
amphibians39, whereas the preBo
¨tC would have emerged later, with the
mammalian aspiration pump31,40–42.
Three independent lines of evidence (acute riluzole application,
transverse sections and Egr2 invalidation) argue that the inactiva-
tion of the e-pF oscillator slows down the E15.5 respiratory-like
rhythm. We previously proposed that the abnormally slow breathing
phenotype of Egr2 null mutants at birth was a result of the
suppression of a rhythm-promoting system that we tentatively
located in the caudal pontine reticular formation6. The present
data indicate that the e-pF is a prime contender for setting the
neonatal breathing pace at a normal frequency. Hence, at E15.5 in
the mouse, a RRG with a dual organization emerges, in which the
e-pF can be considered to contribute in that it increases the rhythm
frequency2, whereas the preBo
¨tC can be considered essential in that
it entrains the motor output30.
In addition to its anatomical layout of and its expression of Phox2b,
several functional properties indicate that the e-pF may be considered
as a forerunner of the neonatal pFRG. First, similar to the pFRG at
neonatal stages, the e-pF upregulates the respiratory-like rhythm2and
drives the incorporation of motor bursts in a quantal manner31,41 at
E15.5. Second, the e-pF at E15.5 includes cells receiving a barrage of
chloride-mediated synaptic currents in phase with the RRG rhythmic
bursts. These inputs, excitatory during the E14.5–E15.5 period28,43,
may transiently contribute to the phasing of the e-pF to the preBo
¨tC,
but will cause inhibition during inspiration and post-inhibitory
rebound excitation on later maturation of the chloride gradient43,
two features associated with pFRG neurons2,44. Third, our preliminary
results indicate that the frequency of the e-pF (but not that of the
preBo
¨tC) oscillator is increased by a low pH challenge at E14.5
(Supplementary Fig. 3), suggesting a role in chemosensitivity that is
consistent with the demonstration that Phox2b expressing neurons of
the pFRG are CO2-sensitive in the neonatal rat20. Neurons ofthe pFRG,
unlike those of the e-pF, show a pre-inspiratory pattern of discharge
(from E19–20 onward in the rat)44. Even though the e-pF cells can also
be described as being pre-active at the earliest stages (E14.5), when their
associated calcium changes precede those in facial motoneurons, this
delay progressively disappears within 24 h as the e-pF and preBo
¨tC
oscillators couple with one another. Thereafter (at E15.5–16.5), the
rhythmic activities of the e-pF, the preBo
¨tC and the motor nerves are
synchronous. Thus, by E15.5 (that is, the time when it begins pacing
fetal breathing), the RRG produces a respiratory-like rhythm charac-
terized by a single inspiratory-like phase28,45,46. How pre-inspiratory
depolarization later arises in e-pF cells, thereby transforming them into
pFRG neurons, remains obscure, as does the location of the presynaptic
inhibitory interneurons. Nonetheless, these data argue that the e-pF
oscillator is the forerunner of the neonatal pFRG.
A reasonable assumption is that the neonatal pFRG may evolve into
the adult RTN20,47. In this context, the e-pF oscillator could represent
yet another, earlier developmental stage of this same entity. Pacemaking
mechanisms relying on INaP and modulation by Ih, present in the e-pF,
may exist only transiently during the postnatal maturation of rhythmic
neural circuits24. Their partial downregulation in cells destined to form
the adult RTN may account for the advent of their characteristic tonic
firing mode18,19.Egr2 null mutants were reported to preserve a
ventilatory response to hypercapnia at birth6. In contrast, the Phox2-
b27Ala/+ mutant, a mouse model for CCHS, has both a severe and
selective loss of Phox2b/NK1R double-positive neurons at E15.5 in a
region encompassing the e-pF oscillator and a complete lack of
sensitivity to hypercapnia16. These findings, together with ours,
would be best explained if the Egr2-dependent, rhythm-promoting
e-pF represented a subset of a larger chemosensitive Phox2b-positive
population. Examining the contribution to the RTN of progenitor
domains producing Phox2b interneurons48,49 outside of Krox20-
expressing rhombomeres should allow for a better understanding of
this potential heterogeneity.
In conclusion, we propose that a two-step developmental process
establishes fetal breathing in mice. In the first step, an Egr2-dependent
e-pF neuronal population clusters at the ventral surface of the
1034 VOLUME 12
[
NUMBER 8
[
AUGUST 2009 NATURE NEUROSCIENCE
ARTICLES
© 2009 Nature America, Inc. All rights reserved.
hindbrain and functions as an oscillator. In a second step, at around
E15.5, a second oscillator, the preBo
¨tC, emerges independently and the
e-pF couples with and entrains it. In this manner, the dual organization
of the RRG is established at the time of inception of fetal breathing.
Early e-pF rhythms preceding the first breathing movements appear
to be of clinical, developmental and evolutionary relevance, and are
required for the respiratory rhythm generator to pace breathing at a
normal frequency during the perinatal period. The potential implica-
tion of this early stage of respiratory development in breathing
disorders, particularly in CCHS, should be considered.
METHODS
Methods and any associated references are available in the online
version of the paper at http://www.nature.com/natureneuroscience/.
Note: Supplementary information is available on the Nature Neuroscience website.
ACKNOWLEDGMENTS
We thank J.F. Brunet and C. Goridis for comments on the manuscript and the
Phox2b antibody, and K. Kullander for VGlut2 mutant mice. This work benefited
from the facilities and expertise of the Imagif Cell Biology Unit of the Gif
campus. This work was supported by the Centre National de la Recherche
Scientifique,Institut National de la Sante
´et de la Recherche Me
´dicale (M.T.-B.)
and ANR grant ANR-07-Neuro-007-01 (G.F.).
AUTHOR CONTRIBUTIONS
M.T.-B. and G.F. designed, performed research and analyzed data, M.K. carried
out immunostaining on Egr2 mutants and N.W. performed pharmacological
treatments at E15.5. J.C. and P.C. helped with interpreting results and G.F. wrote
the manuscript.
Published online at http://www.nature.com/natureneuroscience/.
Reprints and permissions information is available online at http://www.nature.com/
reprintsandpermissions/.
1. Smith, J.C., Ellenberger, H.H., Ballanyi,K., Richter, D.W. & Feldman, J.L.Pre-Botzinger
complex: a brainstem region that may generate respiratory rhythm in mammals. Science
254, 726–729 (1991).
2. Onimaru, H. & Homma, I. A novel functional neuron group for respiratory rhythm
generation in the ventral medulla. J. Neurosci. 23, 1478–1486 (2 003).
3. Blanchi, B. et al. MafB deficiency causes defective respiratory rhythmogenesis and fatal
central apnea at birth. Nat. Neurosci. 6, 1091–1100 (2003).
4. Cheng, L. et al. Tlx3 and Tlx1 are post-mitotic selector genes determining glutamatergic
over GABAergic cell fates. Nat. Neurosci. 7, 510–517 ( 2004).
5. del Toro, E.D. et al. Generation of a novel functional neuronal circuit in Hoxa1 mutant
mice. J. Neurosci. 21, 5637–5642 (2001).
6. Jacquin, T.D. et al. Reorganization of pontine rhythmogenic neuronal networks in Krox-
20 knock-out mice. Neuron 17, 747–758 (1996).
7. Shirasawa, S. et al. Rnx deficiency results in congenital central hypoventilation. Nat.
Genet. 24, 287–290 (2000).
8. Schneider-Maunoury, S. et al. Disruptionof Krox-20 results in alteration of rhombomeres
3 and 5 in the developing hindbrain. Cell 75, 1199–1214 (1993).
9. Wilkinson, D.G., Bhatt, S., Chavrier, P., Bravo, R. & Charnay, P. Segment-specific
expression of a zinc-finger gene in the developing nervous system of the mouse. Nature
337, 461–464 (1989).
10. Coutinho, A.P. et al. Induction of a parafacial rhythm generator by rhombomere 3 in the
chick embryo. J. Neurosci. 24, 9383–9390 (2004).
11. Brunet, J.F. a.G.C. Phox2b and the homeostatic brain. in Genetic Basis f or Respiratory
Control Disorders (Gauthier, C., ed.) 25 (Springer, New York, 2008).
12. Dauger, S. et al. Phox2b controls the development of peripheral chemoreceptors and
afferent visceral pathways. Development 130, 6635–6642 (2003 ).
13. Amiel, J. et al. Polyalanine expansion and frameshift mutations of the paired-like
homeobox gene PHOX2B in congenital central hypoventilation syndrome. Nat. Genet.
33, 459–461 (2003).
14. Weese-Mayer, D.E., Berry-Kravis, E.M. & Marazita, M.L. In pursuit (and discovery) of a
genetic basis for congenital central hypoventilation syndrome. Respir. Physiol. Neuro-
biol. 149, 73–82 (2005).
15. Spengler, C.M., Gozal, D. & Shea, S.A. Chemoreceptive mechanisms elucidated by
studies of congenital central hypoventilation syndrome. Respir. Physiol. 129, 247–255
(2001).
16. Dubreuil, V. et al. A human mutation in Phox2b causes lack of CO2 chemosensitivity,
fatal central apnea and specific loss of parafacial neurons. Proc. Natl. Acad. Sci. USA
105, 1067–1072 (2008).
17. Smith, J.C., Morrison, D.E., Ellenberger, H.H., Otto, M.R. & Feldman, J.L. Brainstem
projections to the major respiratory neuron populations in the medulla of the cat.
J. Comp. Neurol. 281, 69–96 (1989).
18. Mulkey, D.K. et al. Respiratory control by ventral surface chemoreceptor neuronsin rats.
Nat. Neurosci. 7, 1360–1369 (2004).
19. Stornetta, R.L. et al. Expression of Phox2b by brainstem neurons involved in chemo-
sensory integration in the adult rat. J. Neurosci. 26, 10305–10314 (2006).
20. Onimaru, H., Ikeda, K. & Kawakami, K. CO2-sensitive preinspiratory neur ons of the
parafacial respiratory group express Phox2b in the neonatal rat. J. Neurosci. 28,
12845–12850 (2008).
21. Voiculescu, O., Charnay, P. & Schneider-Maunoury, S. Expression pattern of a Krox-20/
Cre knock-in allele in the developing hindbrain, bones, and peripheral nervous system.
Genesis 26, 123–126 (2000).
22. Srinivas, S. et al. Cre reporter strains produced by targeted insertion of EYFP and ECFP
into the ROSA26 locus. BMC Dev. Biol. 1, 4 (2001).
23. Suzue, T. Respiratory rhythm generation in the in vitro brain stem-spinal cord prepara-
tion of the neonatal rat. J. Physiol. (Lond.) 354, 173–183 (1984).
24. Chan, C.S. et al. ‘Rejuvenation’ protects neurons in mouse models of Parkinson’s
disease. Nature 447, 1081–1086 (2007).
25. Walle
´n-Mackenzie, A. et al. Vesicular glutamate transporter 2 is required for central
respiratory rhythm generation, but not for locomotor central pattern generation.
J. Neurosci. 26, 12294–12307 (2006).
26. Anselmi, F. et al. ATP release through connexin hemichannel s and gap junction transfer
of second messengers propagateCa2+ signals across the inner ear. Proc. Natl. Acad.Sci.
USA 105, 18770–18775 (2008).
27. Contreras, J.E., Saez, J.C., Bukauskas, F.F. & Bennett, M.V. Gating and regulation of
connexin 43 (Cx43) hemichannels. Proc. Natl. Acad. Sci. USA 100, 11388–11393
(2003).
28. Thoby-Brisson, M., Trinh, J.B., Champagnat, J. & Fortin, G. Emergence of the pre-
Botzinger respiratory rhythm generator in the mouse embryo. J. Neurosci. 25,
4307–4318 (2005).
29. Gray, P.A., Rekling, J.C., Bocchiaro, C.M. & Feldman, J.L. Modulation of respiratory
frequency by peptidergic input to rhythmogenic neurons in the preBotzinger complex.
Science 286, 1566–1568 (1999).
30. Feldman, J.L. & Del Negro, C.A. Looking forinspiration: new perspectiveson respiratory
rhythm. Nat. Rev. Neurosci. 7, 232–242 (2006).
31. Mellen, N.M., Janczewski, W.A., Bocchiaro, C.M. & Feldman, J.L. Opioid-induced
quantal slowing reveals dual networks for respiratory rhythm generation. Neuron 37,
821–826 (2003).
32. Brunet, J.F. & Goridis, C. Phox2b and the homeostatic brain. in Genetic Basis for
Respiratory Control Disorders (Gauthier, C., ed.) 25–44 (Springer, New York, 2008).
33. Greer, J.J., Smith, J.C.& Feldman, J.L. Role of excitatory amino acids in the generation
and transmission of respiratory drive in neonatal rat. J. Physiol. (Lond.) 437, 727–749
(1991).
34. Rubin, J.E.H.J., Mendenhall, J.L. & Del Negro, C.A. Calcium-activated nonspecific
cation current and synaptic depression promote network-dependent burst oscillations.
Proc. Natl. Acad. Sci. USA 106, 2939–2944 (2009).
35. Fre
`re, S.G., Kuisle, M. & Luthi, A. Regulation of recombinant and native hyperpolariza-
tion-activated cation channels. Mol. Neurobiol. 30, 279–305 (2004).
36. Koizumi,H. et al. Functional imaging,spatial reconstruction,and biophysical analysisof
a respiratory motor circuit isolated in vitro.J. Neurosci. 28, 2353–2365 (2008).
37. Koshiya, N. & Smith, J.C. Neuronal pacemaker for breathing visualized in vitro.Nature
400, 360–363 (1999).
38. Borday, C., Coutinho, A., Germon, I., Champagnat, J. & Fortin, G. Pre-/post-otic
rhombomeric interactions control the emergence of a fetal-like respiratory rhythm in
the mouse embryo. J. Neurobiol. 66, 1285–1301 (2006).
39. Vasilakos, K., Wilson, R.J.,Kimura, N. & Remmers, J.E. Ancient gill andlung oscillators
may generate the respiratory rhythm of frogs and rats. J. Neurobiol. 62, 369–385
(2005).
40. Feldman, J.L., Mitchell, G.S. & Nattie, E.E. Breathing: rhythmicity, plasticity, chemo-
sensitivity. Annu. Rev. Neurosci. 26, 239–266 (2003).
41. Janczewski, W.A. & Feldman, J.L. Distinct rhythm generators for inspiration and
expiration in the juvenile rat. J. Physiol. (Lond.) 570, 407–420 (2006).
42. Milsom, W.K. Evolutionary trends inrespiratory mechanisms. Adv. Exp. Med. Biol. 605,
293–298 (2008).
43. Ren, J. & Greer, J.J. Modulation of respiratory rhythmogenesis by chloride-mediated
conductances during the perinatal period. J. Neurosci. 26, 3721–3730 (2006).
44. Onimaru, H. & Homma, I. Developmental changes in the spatio-temporal pattern of
respiratory neuron activity in the medulla of latefetal rat. Neuroscience 131, 969–977
(2005).
45. Kobayashi, K., Lemke, R.P. & Greer, J.J. Ultrasound measurements of fetal breathing
movements in the rat. J. Appl. Physiol. 91, 316–320 (2001).
46. Pagliardini, S., Ren, J. & Greer, J.J. Ontogeny of the pre-Botzinger complex in perinatal
rats. J. Neurosci. 23, 9575–9584 (2003).
47. Guyenet, P.G. The 2008 Carl LudwigLecture: retrotrapezoid nucleus, CO2homeostasis,
and breathing automaticity. J. Appl. Physiol. 105, 404–416 (2008).
48. Pattyn, A., Morin, X., Cremer, H., Goridis, C. & Brunet, J.-F. Expression and interactions
of the two closely related homeobox genes Phox2a and Phox2b during neurogenesis.
Development 124, 4065–4075 (1997).
49. Pagliardini, S. et al. Central respiratory rhythmogenesis is abnormal in lbx1-deficient
mice. J. Neurosci. 28, 11030–11041 (2008).
NATURE NEUROSCIENCE VOLUME 12
[
NUMBER 8
[
AUGUST 2009 1035
ARTICLES
© 2009 Nature America, Inc. All rights reserved.
ONLINE METHODS
Mouse l ines. All of the mouse lines used in this study were maintained in a
mixed C57Bl6/DBA2 background. The Egr2lacZ allele carries an in-frame inser-
tion of the lacZ coding sequence in the second exon of Egr2 (ref. 50). In the
Egr2cre/+ allele, the Egr2 coding sequence was substituted by the Cre recombi-
nase coding sequence. The R26R-EYFP mouse line22, which allows Cre-
mediated activation of EYFP expression, was kindly provided by F. Costantini
(Columbia University). The day of the vaginal plug was considered E0.5. All
experiments were carried out in accordance with National (JO 87–848) and
European legislation (86/609/CEE) on animal experimentation.
Hindbrain electrophysiology. Pregnant mice were killed by cervical dislocation
at E14.5–E16.5. Embryos were excised from the uterus and kept in oxygenated
artificial cerebrospinal fluid (aCSF) at 20 1C until they were used in electro-
physiological and optical recordings sessions. aCSF was composed of 120 mM
NaCl, 8 mM KCl, 1.26 mM CaCl2, 1.5 mM MgCl2, 21 mM NaHCO3,0.58mM
Na2HPO4and 30 mM glucose (pH ¼7.4). For low pH aCSF, the NaHCO3
concentration was decreased to 10.5 mM and the NaCl concentration was
increased to 130.5 mM. The pH of the superfusate was measured continuously
in the recording chamber with a microelectrode (MI-410, Microelectrodes)
calibrated with standard buffers. A high external [K+] was purposefully used
to ensure maintenance of the functional mode of the preBo
¨tC, as previously
described28, at the time of its emergence to optimally detect early network
interactions modulating the preBo
¨tC activity, although this may have generally
increased neuronal baseline excitability.
En bloc hindbrain preparations were prepared as described previously28 and
were positioned with the ventral side up in the recording chamber. Transverse
450-mm-thick slices were obtained at the level of the nVII or at the level of
the preBo
¨tC28 using a vibratome and were transferred into a 1-ml recording
chamber that was continuously superfused at 2 ml min1with oxygenated
aCSF at 30 1C. The caudal limit of the facial motor nucleus was used as an
anatomical landmark to generate transverse facial slices and preBo
¨tC slices.
To obtain transverse facial slices, we generated a series of 150-mm-thick slices
progressing from caudal to rostral up to the first planes where the presence of the
facial nucleus could be unambiguously distinguished as a result of its higher
optical refringence compared with that of the neighboring tissue. At this point, a
single 450-mm-transverse slice was cut that had an anterior limit corresponding
approximately to the equatorial transverse plan of the facial motor nucleus.
Facial slices were transferred to the recording chamber and positioned with
the anterior side up. For preBo
¨tC slices, the slicing was carried out as described
above, but progressing from rostral to caudal, up to the last plane of the facial
nucleus. At this point, a 250-mm-thickslicewasmadetoreachtheanteriorlimit
of the preBo
¨tC, and a single 450-mm-thick transverse preBo
¨tc slice was produced.
Hypoglossal nerve root activity and population activity (e-pF or preBo
¨tC)
in whole hindbrain preparations were recorded using glass micropipettes
suction electrodes (150-mm tip diameter). For e-pF recordings in whole hind-
brain preparations, the electrodes were positioned on the brain surface, and
for the recording of the preBo
¨tC, the electrodes were progressively inserted at
depths of about 100–150 mm below the hindbrain surface, sufficient to record
the activity of the preBo
¨tC. The micropipettes filled with aCSF were connected
through silver wires to a high-gain alternating current amplifier (Grass, 7P511),
filtered (bandwidth, 3 Hz through 3 kHz), integrated using an electronic filter
(Neurolog System, time constant of 100 ms), recorded on a computer via a
digitizing interface (Digidata 1322A, Molecular Devices) and analyzed with the
pClamp9 software (Molecular Devices). Whole-cell patch-clamp neuronal
recordings were performed under visual control using differential interference
contrast (DIC) and infrared video microscopy, an Axoclamp2A amplifier
(Molecular Devices), a digitizing interface (Digidata 1322A, Molecular Devices)
and the software program pClamp9 (Molecular Devices). Patch electrodes
(resistance of 4–6 MO) were pulled from borosilicate glass tubes (Clark GC
150TF) and filled with a solution containing 140 mM potassium gluconic acid,
1mMCaCl
2,6mMH
2O, 10 mM EGTA, 2 mM MgCl2,4mMNa
2ATP a nd
10 mM HEPES (pH 7.2). In voltage-clamp mode, we analyzed the persistent
sodium current INaP and the hyperpolarization-activated current Ih.INaP was
activated using a slow depolarizing ramp from 60 mV to +10 mV and blocked
by 5–20 mMriluzole(Fig. 2d). The Ihcurrent was evoked by applying
hyperpolarizing voltage steps (from 50 mV to 120 mV) and was blocked
by 100 mM ZD7288. The Ihcurrent/voltage curve was built by measuring the
difference between current amplitudes values measured at the beginning and
the end of each voltage step (Fig. 2f). In current clamp, the Ihcurrent activation
led to depolarizing sags in response to hyperpolarizing current pulses (Fig. 2f).
Drugs were obtained from Sigma, dissolved in aCSF and bath-applied for
10–15 min a final concentration of 0.1 mM Substance P (SP), 0.3 mM DAMGO,
10 mM CNQX, 10 mMAP5,10mMbicuculline,5mM strychnine, 5–20 mM
Riluzole), 50 mMCBX,100mM ZD7288 and 100 mM lanthanum. To minimize
the risk of nonspecific effects resulting from long-term exposure to riluzole,
ZD7288, La3+ and CBX, we measured their effects during a 2-min period
beginning 5 min after switching on the perfusion source containing the tested
compounds, a delay that is approximately tenfold larger than the time constant
of concentration change kinetics of the chamber. In addition, when examining
rhythm generation in the e-pF at E14.5, nonspecific effects of these compounds
that were linked to interactions with glutamatergic transmission, being dis-
pensable, were probably minimal. We examined whether (e-pF cell recordings,
n¼4, data not shown) 50 mM CBX had an effect on the amplitude and kinetics
of action potential and on the kinetics of synaptic currents. Values are given as
means ± s.e.m. Differences were regarded as significant at Po0.05.
Calcium imaging. Whole hindbrain and slices were incubated for 40 min in
oxygenated aCSF containing the cell-permeable calcium indicator dye Calcium-
Green 1AM (10 mM, Molecular Probes). Whole hindbrain preparations were
positioned in the recording chamber with the ventral side up. After a 30-min
recovery period in the recording chamber to wash out the dye excess, a
standard epi-fluorescent illumination system on an E-600-FN upright micro-
scope (Nikon) equipped with a fluorescein filter block was used to excite the
dye and capture the emitted light. Fluorescence images were captured with a
cooled CCD camera (Coolsnap HQ, Photometrics) using an exposure time of
100 ms in overlapping mode (simultaneous exposure and readout) during
periods of 60–180 s and analyzed using Metamorph software (Universal
Imaging). To perform calcium imaging of YFP-expressing cells, we first
acquired images of EYFP cells with corresponding DIC images. After dye
loading, a careful positioning over the same cellular field was ascertained
through visualizing cellular profiles using DIC images and adjusting their
alignments. The EYFP-labeled cell false-colored red image and the calcium-
loaded cell false-colored green image were overlaid (Fig. 1g)todetermine
double-labeled somas and to position regions of interest for measurements of
fluorescence changes (Fig. 1g). In all cases, the average intensity in a region of
interest was calculated for each frame and the changes in fluorescence were
normalized to their initial value by expression as the ratio of changes in
fluorescence to initial fluorescence (F/F).
Immunofluorescence. Mouse embryos were fixed for 2–3 h in 4% para-
formaldehyde (wt/vol) in phosphate-saline buffer, cryoprotected in 25% sucrose
(wt/vol), embedded in O.C.T. Compound (Tissue-Tek) and sectioned at 14 or
20 mm. Immunohistochemistry was performed on frozen sections as previously
described28. We used antibodies to rabbit NK1R (Sigma, 1:5,000), guinea pig
Islet1/2 (gift from J. Ericson, Karolinska Institute, 1:1500), chick GFP (Aves Lab,
1:2,000) and rabbit Phox2b (gift from C. Goridis, ENS, 1:1,500). Species
specific antibodies conjugated to Alexa 488, (Molecular Probes), Cy3 and
Cy5 (Jackson ImmunoResearch) were used. Biocytin-filled neurons were
labeled using Extra-Avidin-FITC (Sigma, 1:400). Rocking incubation with
primary antibodies was carried out overnight at 4 1C and secondary antibodies
were incubated for 3 h. The material was mounted in Vectashield medium
(Vector Labs). Counts of e-pF cells were made in an area delimited ventrally by
the medullary surface, dorsally by the nVII, rostrally by the rostral end of the
nVII and extending 100 mm caudal to the nVII. Cells were counted in the
transverse plan using all consecutive 20-mmsections(n¼2) or every other
section (n¼1) and in the sagittal plane on every other section (n¼1). In all
cases, cells were counted on both sides; estimation of the total number of cells
included a multiplication factor of 2 when appropriate. Fluorescent labeling
was visualized on a Leica SP2 confocal microscope. All figures were color
corrected and assembled using Adobe Photoshop and Illustrator.
50. Voiculescu, O. et al. Hindbrain patterning: Krox20 couples segmentation and specifica-
tion of regional identity. Development 128, 4967–4978 (2001).
NATURE NEUROSCIENCE doi:10.1038/nn.2354
© 2009 Nature America, Inc. All rights reserved.
... The voltage-dependent fast sodium channels and persistent sodium channels are involved in respiratory burst generation [13,15,17,21,27]. Moreover, sodium leak channels can affect the excitability of neurons in the respiratory center by influencing the resting membrane potential [24,30]. ...
Article
Full-text available
Aconitine is a sodium channel opener, but its effects on the respiratory center are not well understood. We investigated the dose-dependent effects of aconitine on central respiratory activity in brainstem-spinal cord preparations isolated from newborn rats. Bath application of 0.5–5 μM aconitine caused an increase in respiratory rhythm and decrease in the inspiratory burst amplitude of the fourth cervical ventral root (C4). Separate application of aconitine revealed that medullary neurons were responsible for the respiratory rhythm increase, and neurons in both the medulla and spinal cord were involved in the decrease of C4 amplitude by aconitine. A local anesthetic, lidocaine (100 μM), or a voltage-dependent sodium channel blocker, tetrodotoxin (0.1 μM), partially antagonized the C4 amplitude decrease by aconitine. Tetrodotoxin treatment tentatively decreased the respiratory rhythm, but lidocaine tended to further increase the rhythm. Treatment with 100 μM riluzole or 100 μM flufenamic acid, which are known to inhibit respiratory pacemaker activity, did not reduce the respiratory rhythm enhanced by aconitine + lidocaine. The application of 1 μM aconitine depolarized the preinspiratory, expiratory, and inspiratory motor neurons. The facilitated burst rhythm of inspiratory neurons after aconitine disappeared in a low Ca2+/high Mg2+ synaptic blockade solution. We showed the dose-dependent effects of aconitine on respiratory activity. The antagonists reversed the depressive effects of aconitine in different manners, possibly due to their actions on different sites of sodium channels. The burst-generating pacemaker properties of neurons may not be involved in the generation of the facilitated rhythm after aconitine treatment.
... RTN neurons are intrinsically sensitive to changes in CO 2 /H + across a variety of in vitro preparations, including brainstem-spinal cord preparations, acute or cultured brainstem slices and, importantly, acutely dissociated neurons ( Figure 2A) (Mulkey et al., 2006;Mulkey et al., 2007a;Lazarenko et al., 2010b;Hawryluk et al., 2012;Wenker et al., 2012;Wang et al., 2013a;Sobrinho et al., 2014;Hawkins et al., 2015;Kumar et al., 2015;Wu et al., 2019). During early development, a group of CO 2 /H + sensitive, Phox2b-expressing neurons in the parafacial region display rhythmic pre-and post-inspiratory firing patterns in brainstemspinal cord preparations; these have been called the embryonic parafacial oscillator (ePF) or, in the early postnatal period (P0-P2), the parafacial respiratory group (pFRG), and are most likely early precursors to the RTN (Onimaru et al., 2008;Thoby-Brisson et al., 2009;Ruffault et al., 2015). In slightly older neonatal brainstem slice preparations (>P6), RTN neurons are tonically active at physiological pH levels, depolarize and increase action potential firing during bath acidification, and hyperpolarize and decrease firing during bath alkalization. ...
Article
Full-text available
An interoceptive homeostatic system monitors levels of CO 2 /H ⁺ and provides a proportionate drive to respiratory control networks that adjust lung ventilation to maintain physiologically appropriate levels of CO 2 and rapidly regulate tissue acid-base balance. It has long been suspected that the sensory cells responsible for the major CNS contribution to this so-called respiratory CO 2 /H ⁺ chemoreception are located in the brainstem—but there is still substantial debate in the field as to which specific cells subserve the sensory function. Indeed, at the present time, several cell types have been championed as potential respiratory chemoreceptors, including neurons and astrocytes. In this review, we advance a set of criteria that are necessary and sufficient for definitive acceptance of any cell type as a respiratory chemoreceptor. We examine the extant evidence supporting consideration of the different putative chemoreceptor candidate cell types in the context of these criteria and also note for each where the criteria have not yet been fulfilled. By enumerating these specific criteria we hope to provide a useful heuristic that can be employed both to evaluate the various existing respiratory chemoreceptor candidates, and also to focus effort on specific experimental tests that can satisfy the remaining requirements for definitive acceptance.
... In embryonic and newborn rodents, the lateral parafacial nucleus is rhythmically active at rest (Onimaru and Homma, 2003;Thoby-Brisson et al., 2009), but this activity wanes during early development (Oku et al., 2007;van der Heijden and Zoghbi, 2020), and in adults the lateral parafacial nucleus is generally silent during eupnea, but rhythmically active during hyperpnea (Pagliardini et al., 2011;Huckstepp et al., 2015;de Britto and Moraes, 2017;Figure 3C-D). The inactivity at rest could be due to tonic inhibition from the medial NTS . ...
Article
Full-text available
Respiration is a brain function on which our lives essentially depend. Control of respiration ensures that the frequency and depth of breathing adapt continuously to metabolic needs. In addition, the respiratory control network of the brain has to organize muscular synergies that integrate ventilation with posture and body movement. Finally, respiration is coupled to cardiovascular function and emotion. Here, we argue that the brain can handle this all by integrating a brainstem central pattern generator circuit in a larger network that also comprises the cerebellum. Although currently not generally recognized as a respiratory control center, the cerebellum is well known for its coordinating and modulating role in motor behavior, as well as for its role in the autonomic nervous system. In this review, we discuss the role of brain regions involved in the control of respiration, and their anatomical and functional interactions. We discuss how sensory feedback can result in adaptation of respiration, and how these mechanisms can be compromised by various neurological and psychological disorders. Finally, we demonstrate how the respiratory pattern generators are part of a larger and integrated network of respiratory brain regions.
... The respiratory rhythmicity arises from the coordinated interaction of two anatomically and functionally distinct oscillators in the brainstem: an inspiratory oscillator at the preBötzinger complex (preBötC), the kernel of respiration that drives inspiratory musculature (Smith et al., 1991;Tan et al., 2008), and a conditional expiratory oscillator in the lateral parafacial group (pF L ; also termed parafacial respiratory group, pFRG), driving active expiration in states of elevated metabolic demand (Mellen et al., 2003;Onimaru and Homma, 2003;Janczewski and Feldman, 2006;Abdala et al., 2009;Huckstepp et al., 2015). This two-oscillator system establishes during early embryonic development, varying its hierarchical organization during pre-and postnatal maturation and the degree of activation of each one with the metabolic and physiological states (e.g., rest/exercise or sleep/wake) (Oku et al., 2007;Thoby-Brisson et al., 2009;Pagliardini et al., 2011;Andrews and Pagliardini, 2015;Huckstepp et al., 2016;Leirão et al., 2018). However, a fundamental unresolved question is how this twooscillator system can generate periodic breathing. ...
Article
Full-text available
Periodic Cheyne–Stokes breathing (CSB) oscillating between apnea and crescendo–decrescendo hyperpnea is the most common central apnea. Currently, there is no proven therapy for CSB, probably because the fundamental pathophysiological question of how the respiratory center generates this form of breathing instability is still unresolved. Therefore, we aimed to determine the respiratory motor pattern of CSB resulting from the interaction of inspiratory and expiratory oscillators and identify the neural mechanism responsible for breathing regularization induced by the supplemental CO2 administration. Analysis of the inspiratory and expiratory motor pattern in a transgenic mouse model lacking connexin-36 electrical synapses, the neonatal (P14) Cx36 knockout male mouse, with a persistent CSB, revealed that the reconfigurations recurrent between apnea and hyperpnea and vice versa result from cyclical turn on/off of active expiration driven by the expiratory oscillator, which acts as a master pacemaker of respiration and entrains the inspiratory oscillator to restore ventilation. The results also showed that the suppression of CSB by supplemental 12% CO2 in inhaled air is due to the stabilization of coupling between expiratory and inspiratory oscillators, which causes the regularization of respiration. CSB rebooted after washout of CO2 excess when the inspiratory activity depressed again profoundly, indicating that the disability of the inspiratory oscillator to sustain ventilation is the triggering factor of CSB. Under these circumstances, the expiratory oscillator activated by the cyclic increase of CO2 behaves as an “anti-apnea” center generating the crescendo–decrescendo hyperpnea and periodic breathing. The neurogenic mechanism of CSB identified highlights the plasticity of the two-oscillator system in the neural control of respiration and provides a rationale base for CO2 therapy.
... A subset of neurons within the preBötC co-express the neurokinin 1 receptor (NK1R) and somatostatin (SST) [50]. Phox2b is expressed in the parafacial respiratory group within the ventrolateral medulla [51] and, hence, serves as a marker to distinguish the neurons of the preBötC from neighboring neuronal subpopulations ( Figure S1) in tissue sections from fixed mouse brains at E18.5. ...
Article
Full-text available
KCC2 mediates extrusion of K+ and Cl− and assuresthe developmental “switch” in GABA function during neuronal maturation. However, the molecular mechanisms underlying KCC2 regulation are not fully elucidated. We investigated the impact of transforming growth factor beta 2 (TGF-β2) on KCC2 during neuronal maturation using quantitative RT-PCR, immunoblotting, immunofluorescence and chromatin immunoprecipitation in primary mouse hippocampal neurons and brain tissue from Tgf-β2-deficient mice. Inhibition of TGF-β/activin signaling downregulates Kcc2 transcript in immature neurons. In the forebrain of Tgf-β2−/− mice, expression of Kcc2, transcription factor Ap2β and KCC2 protein is downregulated. AP2β binds to Kcc2 promoter, a binding absent in Tgf-β2−/−. In hindbrain/brainstem tissue of Tgf-β2−/− mice, KCC2 phosphorylation at T1007 is increased and approximately half of pre-Bötzinger-complex neurons lack membrane KCC2phenotypes rescued through exogenous TGF-β2. These results demonstrate that TGF-β2 regulates KCC2 transcription in immature neurons, possibly acting upstream of AP2β, and contributes to the developmental dephosphorylation of KCC2 at T1007. The present work suggests multiple and divergent roles for TGF-β2 on KCC2 during neuronal maturation and provides novel mechanistic insights for TGF-β2-mediated regulation of KCC2 gene expression, posttranslational modification and surface expression. We propose TGF-β2 as a major regulator of KCC2 with putative implications for pathophysiological conditions.
... On the other hand, neurons of the parafacial nucleus are located lateral to the facial motor nucleus. They are thought to control active expiration, a particular type of expiration that is produced when high metabolic demands induce an increase in respiration Thoby-Brisson et al., 2009;Bayliss et al., 2015;Huckstepp et al., 2015Huckstepp et al., , 2018Korsak et al., 2018;Pisanski and Pagliardini, 2019). In 2009, Dubreuil et al. (2009 identified that virtually all retrotrapezoid and parafacial neurons originate from Egr2expressing cells (that is from rhombomeres 3 and/or 5; see Figure 2). ...
Article
Full-text available
Breathing (or respiration) is an unconscious and complex motor behavior which neuronal drive emerges from the brainstem. In simplistic terms, respiratory motor activity comprises two phases, inspiration (uptake of oxygen, O 2) and expiration (release of carbon dioxide, CO 2). Breathing is not rigid, but instead highly adaptable to external and internal physiological demands of the organism. The neurons that generate, monitor, and adjust breathing patterns locate to two major brainstem structures, the pons and medulla oblongata. Extensive research over the last three decades has begun to identify the developmental origins of most brainstem neurons that control different aspects of breathing. This research has also elucidated the transcriptional control that secures the specification of brainstem respiratory neurons. In this review, we aim to summarize our current knowledge on the transcriptional regulation that operates during the specification of respiratory neurons, and we will highlight the cell lineages that contribute to the central respiratory circuit. Lastly, we will discuss on genetic disturbances altering transcription factor regulation and their impact in hypoventilation disorders in humans.
... When combined, these markers form a molecular code that can define the RTN: Phox2b + Nmb + Nk1R + Atoh1 L Egr2 L . In fact, this code could even be further developed (22)(23)(24)(25)(26)(27)(28)(29)(30). Indeed, unique or restricted markers like Phox2b and Nmb have proven their worth in numerous studies elucidating the role of the RTN in breathing behavior. ...
Article
Full-text available
The rhythmicity of breath is vital for normal physiology. Even so, breathing is enriched with multifunctionality. External signals constantly change breathing, stopping it when under water or deepening it during exertion. Internal cues utilize breath to express emotions such as sighs of frustration and yawns of boredom. Breathing harmonizes with other actions that use our mouth and throat, including speech, chewing, and swallowing. In addition, our perception of breathing intensity can dictate how we feel, such as during the slow breathing of calming meditation and anxiety-inducing hyperventilation. Heartbeat originates from a peripheral pacemaker in the heart, but the automation of breathing arises from neural clusters within the brainstem, enabling interaction with other brain areas and thus multifunctionality. Here, we document how the recent transformation of cellular and molecular tools has contributed to our appreciation of the diversity of neuronal types in the breathing control circuit and how they confer the multifunctionality of breathing. Expected final online publication date for the Annual Review of Physiology, Volume 85 is February 2023. Please see http://www.annualreviews.org/page/journal/pubdates for revised estimates.
Article
Opioid overdose is the leading cause of drug overdose lethality, posing an urgent need for investigation. The key brain region for inspiratory rhythm regulation and opioid-induced respiratory depression (OIRD) is the preBötzinger Complex (preBötC) and current knowledge has mainly been obtained from animal systems. This study aims to establish a protocol to generate human preBötC neurons from induced pluripotent cells (iPSCs) and develop an opioid overdose and recovery model utilizing these iPSC-preBötC neurons. A de novo protocol to differentiate preBötC-like neurons from human iPSCs is established. These neurons express essential preBötC markers analyzed by immunocytochemistry and demonstrate expected electrophysiological responses to preBötC modulators analyzed by patch clamp electrophysiology. The correlation of the specific biomarkers and function analysis strongly suggests a preBötC-like phenotype. Moreover, the dose-dependent inhibition of these neurons' activity is demonstrated for four different opioids with identified IC50's comparable to the literature. Inhibition is rescued by naloxone in a concentration-dependent manner. This iPSC-preBötC mimic is crucial for investigating OIRD and combating the overdose crisis and a first step for the integration of a functional overdose model into microphysiological systems.
Article
Microglia, brain-resident macrophages, are key players in brain development, regulating synapse density, shaping neural circuits, contributing to plasticity, and maintaining nervous tissue homeostasis. These functions are ensured from early prenatal development until maturity, in normal and pathological states of the central nervous system. Microglia dysfunction can be involved in several neurodevelopmental disorders, some of which are associated with respiratory deficits. Breathing is a rhythmic motor behavior generated and controlled by hindbrain neuronal networks. The operation of the central respiratory command relies on the proper development of these rhythmogenic networks, formation of their appropriate interactions, and their lifelong constant adaptation to physiological needs. This review, focusing exclusively on the perinatal period, outlines recent advances obtained in rodents in determining the roles of microglia in the establishment and functioning of the respiratory networks and their involvement in certain pathologies.
Article
Full-text available
Neurokinin-1 receptor (NK1R) and μ-opioid receptor (μOR) agonists affected respiratory rhythm when injected directly into the preBötzinger Complex (preBötC), the hypothesized site for respiratory rhythmogenesis in mammals. These effects were mediated by actions on preBötC rhythmogenic neurons. The distribution of NK1R⁺ neurons anatomically defined the preBötC. Type 1 neurons in the preBötC, which have rhythmogenic properties, expressed both NK1Rs and μORs, whereas type 2 neurons expressed only NK1Rs. These findings suggest that the preBötC is a definable anatomic structure with unique physiological function and that a subpopulation of neurons expressing both NK1Rs and μORs generate respiratory rhythm and modulate respiratory frequency.
Article
Full-text available
Several Cre reporter strains of mice have been described, in which a lacZ gene is turned on in cells expressing Cre recombinase, as well as their daughter cells, following Cre-mediated excision of a loxP-flanked transcriptional "stop" sequence. These mice are useful for cell lineage tracing experiments as well as for monitoring the expression of Cre transgenes. The green fluorescent protein (GFP) and variants such as EYFP and ECFP offer an advantage over lacZ as a reporter, in that they can be easily visualized without recourse to the vital substrates required to visualize β-gal in living tissue. In view of the general utility of targeting the ubiquitously expressed ROSA26 locus, we constructed a generic ROSA26 targeting vector. We then generated two reporter lines of mice by inserting EYFP or ECFP cDNAs into the ROSA26 locus, preceded by a loxP-flanked stop sequence. These strains were tested by crossing them with transgenic strains expressing Cre in a ubiquitous (β-actin-Cre) or a cell-specific (Isl1-Cre and En1-Cre) pattern. The resulting EYFP or ECFP expression patterns indicated that the reporter strains function as faithful monitors of Cre activity. In contrast to existing lacZ reporter lines, where lacZ expression cannot easily be detected in living tissue, the EYFP and ECFP reporter strains are useful for monitoring the expression of Cre and tracing the lineage of these cells and their descendants in cultured embryos or organs. The non-overlapping emission spectra of EYFP and ECFP make them ideal for double labeling studies in living tissues.
Article
Full-text available
Central pattern generators (CPGs) produce neural-motor rhythms that often depend on specialized cellular or synaptic properties such as pacemaker neurons or alternating phases of synaptic inhibition. Motivated by experimental evidence suggesting that activity in the mammalian respiratory CPG, the preBötzinger complex, does not require either of these components, we present and analyze a mathematical model demonstrating an unconventional mechanism of rhythm generation in which glutamatergic synapses and the short-term depression of excitatory transmission play key rhythmogenic roles. Recurrent synaptic excitation triggers postsynaptic Ca(2+)-activated nonspecific cation current (I(CAN)) to initiate a network-wide burst. Robust depolarization due to I(CAN) also causes voltage-dependent spike inactivation, which diminishes recurrent excitation and thus attenuates postsynaptic Ca(2+) accumulation. Consequently, activity-dependent outward currents-produced by Na/K ATPase pumps or other ionic mechanisms-can terminate the burst and cause a transient quiescent state in the network. The recovery of sporadic spiking activity rekindles excitatory interactions and initiates a new cycle. Because synaptic inputs gate postsynaptic burst-generating conductances, this rhythm-generating mechanism represents a new paradigm that can be dubbed a 'group pacemaker' in which the basic rhythmogenic unit encompasses a fully interdependent ensemble of synaptic and intrinsic components. This conceptual framework should be considered as an alternative to traditional models when analyzing CPGs for which mechanistic details have not yet been elucidated.
Article
Full-text available
Extracellular ATP controls various signaling systems including propagation of intercellular Ca²⁺ signals (ICS). Connexin hemichannels, P2x7 receptors (P2x7Rs), pannexin channels, anion channels, vesicles, and transporters are putative conduits for ATP release, but their involvement in ICS remains controversial. We investigated ICS in cochlear organotypic cultures, in which ATP acts as an IP3-generating agonist and evokes Ca²⁺ responses that have been linked to noise-induced hearing loss and development of hair cell-afferent synapses. Focal delivery of ATP or photostimulation with caged IP3 elicited Ca²⁺ responses that spread radially to several orders of unstimulated cells. Furthermore, we recorded robust Ca²⁺ signals from an ATP biosensor apposed to supporting cells outside the photostimulated area in WT cultures. ICS propagated normally in cultures lacking either P2x7R or pannexin-1 (Px1), as well as in WT cultures exposed to blockers of anion channels. By contrast, Ca²⁺ responses failed to propagate in cultures with defective expression of connexin 26 (Cx26) or Cx30. A companion paper demonstrates that, if expression of either Cx26 or Cx30 is blocked, expression of the other is markedly down-regulated in the outer sulcus. Lanthanum, a connexin hemichannel blocker that does not affect gap junction (GJ) channels when applied extracellularly, limited the propagation of Ca²⁺ responses to cells adjacent to the photostimulated area. Our results demonstrate that these connexins play a dual crucial role in inner ear Ca²⁺ signaling: as hemichannels, they promote ATP release, sustaining long-range ICS propagation; as GJ channels, they allow diffusion of Ca²⁺-mobilizing second messengers across coupled cells. • deafness • mouse models • P2x7 receptor • pannexin • biosensor cells
Article
Full-text available
Phox2b protein is a specific marker for neurons in the parafacial region of the ventral medulla, which are proposed to play a role in central chemoreception and postnatal survival. Mutations of PHOX2B cause congenital central hypoventilation syndrome. However, there have been no reports concerning electrophysiological characteristics of these Phox2b-expressing neurons in the parafacial region of the neonate immediately after birth. This region overlaps with the parafacial respiratory group (pFRG) composed predominantly of preinspiratory (Pre-I) neurons that are involved in respiratory rhythm generation. We studied (1) whether pFRG neurons are Phox2b immunoreactive or not and (2) whether they show intrinsic CO(2) chemosensitivity. We found that most pFRG/Pre-I neurons were Phox2b immunoreactive and depolarized upon increase in CO(2) concentration under condition of action potential-dependent synaptic transmission blockade by tetrodotoxin. We also confirmed that these pFRG neurons expressed neurokinin-1 receptor. They were tyrosine hydroxylase negative and presumed to be glutamatergic. Our findings suggest that Phox2b-expressing parafacial neurons play a role in respiratory rhythm generation as well as central chemoreception and thus are essential for postnatal survival.
Article
Full-text available
Lbx1 is a transcription factor that determines neuronal cell fate and identity in the developing medulla and spinal cord. Newborn Lbx1 mutant mice die of respiratory distress during the early postnatal period. Using in vitro brainstem-spinal cord preparations we tested the hypothesis that Lbx1 is necessary for the inception, development and modulation of central respiratory rhythmogenesis. The inception of respiratory rhythmogenesis at embryonic day 15 (E15) was not perturbed in Lbx1 mutant mice. However, the typical age-dependent increase in respiratory frequency observed in wild-type from E15 to P0 was not observed in Lbx1 mutant mice. The slow respiratory rhythms in E18.5 Lbx1 mutant preparations were increased to wild-type frequencies by application of substance P, thyrotropin releasing hormone, serotonin, noradrenaline, or the ampakine drug 1-(1,4-benzodioxan-6-yl-carbonyl) piperidine. Those data suggest that respiratory rhythm generation within the pre-Bötzinger complex (preBötC) is presumably functional in Lbx1 mutant mice with additional neurochemical drive. This was supported by anatomical data showing that the gross structure of the preBötC was normal, although there were major defects in neuronal populations that provide important modulatory drive to the preBötC including the retrotrapezoid nucleus, catecholaminergic brainstem nuclei, nucleus of the solitary tract, and populations of inhibitory neurons in the ventrolateral and dorsomedial medullary nuclei. Finally, we determined that those defects were caused by abnormalities of neuronal specification early in development or subsequent neuronal migration.
Article
The genetic basis for the development of brainstem neurons that generate respiratory rhythm is unknown. Here we show that mice deficient for the transcription factor MafB die from central apnea at birth and are defective for respiratory rhythmogenesis in vitro. MafB is expressed in a subpopulation of neurons in the preBötzinger complex (preBötC), a putative principal site of rhythmogenesis. Brainstems from Mafb-/- mice are insensitive to preBötC electrolytic lesion or stimulation and modulation of rhythmogenesis by hypoxia or peptidergic input. Furthermore, in Mafb-/- mice the preBötC, but not major neuromodulatory groups, presents severe anatomical defects with loss of cellularity. Our results show an essential role of MafB in central respiratory control, possibly involving the specification of rhythmogenic preBötC neurons.
Chapter
The nervous system, like any other organ, forms under the control of developmental transcription factors. These transcription factors, expressed in varied combinations, commonly called "transcriptional codes", specify regions of the brain and, subsequently, the multitude of neuronal types arising from them. This extremely combinatorial system entails that the expression pattern of any individual transcription factor in the forming brain has little predictive value for the final wiring and function of its constituent neurons. One transcription factor, however, enigmatically stands out: the paired-like homeobox gene Phox2b, specifically expressed and required in neurons which go on to form the visceral reflex circuits that control the digestive, cardiovascular and respiratory systems. We will describe the system-wide implication of Phox2b in the ontogeny of the visceral nervous system and discuss its embryological, physiopathological and evolutionary ramifications. © 2008 Springer Science+Business Media, LLC. All rights reserved.