The H3K27 Demethylase JMJD3 Is Required for Maintenance of the Embryonic Respiratory Neuronal...

23
Cell Reports Article The H3K27 Demethylase JMJD3 Is Required for Maintenance of the Embryonic Respiratory Neuronal Network, Neonatal Breathing, and Survival Thomas Burgold, 1 Nicolas Voituron, 3,6,7 Marieta Caganova, 2,6 Prem Prakash Tripathi, 1,6 Clement Menuet, 3 Betsabeh Khoramian Tusi, 1,8 Fabio Spreafico, 1 Michelle Be ´ vengut, 3 Christian Gestreau, 3 Serena Buontempo, 1 Antonio Simeone, 4 Laurens Kruidenier, 5 Gioacchino Natoli, 1 Stefano Casola, 2 Ge ´ rard Hilaire, 3 and Giuseppe Testa 1, * 1 European Institute of Oncology (IEO) 2 FIRC Institute of Molecular Oncology Foundation (IFOM) IFOM-IEO Campus, Via Adamello 16, 20139 Milan, Italy 3 Mp3-respiration, Centre de Recherche en Neurobiologie et Neurophysiologie de Marseille, Unite ´ Mixte de Recherche 6132 CNRS-Universite ´ Aix-Marseille II et III, Faculte ´ Saint Je ´ ro ˆ me, 13397 Marseille Cedex 20, France 4 CEINGE Biotecnologie Avanzate, Via Comunale Margherita 482, 80145 and Institute of Genetics and Biophysics ‘‘A. Buzzati-Traverso,’’ National Research Council, Via Pietro Castellino 111, 80131 Naples, Italy 5 Epinova DPU, Immuno-Inflammation Therapy Area, GlaxoSmithKline R&D, Medicines Research Centre, Gunnels Wood Road, Stevenage SG1 2NY, UK 6 These authors contributed equally to this work 7 Present address: Laboratoire Re ´ ponses Cellulaires et Fonctionnelles a ` l’Hypoxie, EA 2363 UFR Sante ´ , Me ´ decine, Biologie Humaine, Universite ´ Paris 13, 74 rue Marcel Cachin, 93017 Bobigny Cedex, France 8 Present address: University of Florida College of Medicine, Gainesville, FL 32611, USA *Correspondence: [email protected] http://dx.doi.org/10.1016/j.celrep.2012.09.013 SUMMARY JMJD3 (KDM6B) antagonizes Polycomb silencing by demethylating lysine 27 on histone H3. The inter- play of methyltransferases and demethylases at this residue is thought to underlie critical cell fate transitions, and the dynamics of H3K27me3 during neurogenesis posited for JMJD3 a critical role in the acquisition of neural fate. Despite evidence of its involvement in early neural commitment, however, its role in the emergence and maturation of the mammalian CNS remains unknown. Here, we inacti- vated Jmjd3 in the mouse and found that its loss causes perinatal lethality with the complete and selective disruption of the pre-Bo ¨ tzinger complex (PBC), the pacemaker of the respiratory rhythm generator. Through genetic and electrophysiological approaches, we show that the enzymatic activity of JMJD3 is selectively required for the maintenance of the PBC and controls critical regulators of PBC activity, uncovering an unanticipated role of this en- zyme in the late structuring and function of neuronal networks. INTRODUCTION Posttranslational modifications of histone tails have been asso- ciated, and at times causally linked, to transcriptional outcomes. Due to its comparative stability over other modifications (Byvoet et al., 1972), lysine methylation was initially proposed as an ideal candidate to mediate the epigenetic maintenance of tran- scriptional states in both replicating and postmitotic cells. In particular, di- and trimethylation of lysine 27 on histone H3 (H3K27me2/3) by EZH2 within Polycomb repressive complex 2 (PRC2) has emerged as a critical mechanism of gene repression and, more recently, as a bona fide epigenetic mark that can be propagated through cell division (Hansen et al., 2008; Margueron et al., 2009). Two lines of experimentation revealed, however, the unexpected dynamics of this mark: (1) the identification of H3K27me2/3-specific JmjC-domain demethylases JMJD3 and UTX (Agger et al., 2007; Burgold et al., 2008; De Santa et al., 2007; Jepsen et al., 2007; Lan et al., 2007; Xiang et al., 2007); and (2) the description of genome-wide changes in H3K27me3 that accompany the entire process of neural fate acquisition (Mohn et al., 2008). The developing CNS has indeed become a paradigm-setting model to interrogate how H3K27me3- mediated repression times the activity of critical regulators of neurogenesis, as shown through the ablation of Ezh2 at different stages of corticogenesis whose divergent phenotypes of delayed or accelerated neurogenesis (Hirabayashi et al., 2009; Pereira et al., 2010) were recently reconciled in a model of Polycomb- mediated developmental timing that reflects both the epigenetic transmission of the mark and its dynamic relocation to new targets (Testa, 2011). Conversely, the exquisite sensitivity of neurogenesis to modulation of H3K27me3 underscores the rele- vance of timed H3K27me3 demethylation, as revealed by the crit- ical role of JMJD3 in both mouse embryonic stem cell (ESC) neurulation (Burgold et al., 2008) and chick spinal cord neurogen- esis (Akizu et al., 2010). Yet, despite these early advances the relevance of JMJD3 for the function of the mammalian CNS in vivo remains to be explored and nothing is known about its role in later stages of CNS development and in neuronal maturation. Cell Reports 2, 1–15, October 25, 2012 ª2012 The Authors 1 Please cite this article in press as: Burgold et al., The H3K27 Demethylase JMJD3 Is Required for Maintenance of the Embryonic Respiratory Neuronal Network, Neonatal Breathing, and Survival, Cell Reports (2012), http://dx.doi.org/10.1016/j.celrep.2012.09.013

Transcript of The H3K27 Demethylase JMJD3 Is Required for Maintenance of the Embryonic Respiratory Neuronal...

Please cite this article in press as: Burgold et al., The H3K27 Demethylase JMJD3 Is Required for Maintenance of the Embryonic Respiratory NeuronalNetwork, Neonatal Breathing, and Survival, Cell Reports (2012), http://dx.doi.org/10.1016/j.celrep.2012.09.013

Cell Reports

Article

The H3K27 Demethylase JMJD3 Is Required forMaintenance of the Embryonic Respiratory NeuronalNetwork, Neonatal Breathing, and SurvivalThomas Burgold,1 Nicolas Voituron,3,6,7 Marieta Caganova,2,6 Prem Prakash Tripathi,1,6 Clement Menuet,3

Betsabeh Khoramian Tusi,1,8 Fabio Spreafico,1 Michelle Bevengut,3 Christian Gestreau,3 Serena Buontempo,1

Antonio Simeone,4 Laurens Kruidenier,5 Gioacchino Natoli,1 Stefano Casola,2 Gerard Hilaire,3 and Giuseppe Testa1,*1European Institute of Oncology (IEO)2FIRC Institute of Molecular Oncology Foundation (IFOM)

IFOM-IEO Campus, Via Adamello 16, 20139 Milan, Italy3Mp3-respiration, Centre de Recherche en Neurobiologie et Neurophysiologie deMarseille, Unite Mixte de Recherche 6132CNRS-Universite

Aix-Marseille II et III, Faculte Saint Jerome, 13397 Marseille Cedex 20, France4CEINGE Biotecnologie Avanzate, Via Comunale Margherita 482, 80145 and Institute of Genetics and Biophysics ‘‘A. Buzzati-Traverso,’’National Research Council, Via Pietro Castellino 111, 80131 Naples, Italy5Epinova DPU, Immuno-Inflammation Therapy Area, GlaxoSmithKline R&D, Medicines Research Centre, Gunnels Wood Road,

Stevenage SG1 2NY, UK6These authors contributed equally to this work7Present address: Laboratoire Reponses Cellulaires et Fonctionnelles a l’Hypoxie, EA 2363 UFR Sante, Medecine, Biologie Humaine,

Universite Paris 13, 74 rue Marcel Cachin, 93017 Bobigny Cedex, France8Present address: University of Florida College of Medicine, Gainesville, FL 32611, USA

*Correspondence: [email protected]://dx.doi.org/10.1016/j.celrep.2012.09.013

SUMMARY

JMJD3 (KDM6B) antagonizes Polycomb silencingby demethylating lysine 27 on histone H3. The inter-play of methyltransferases and demethylases atthis residue is thought to underlie critical cell fatetransitions, and the dynamics of H3K27me3 duringneurogenesis posited for JMJD3 a critical role inthe acquisition of neural fate. Despite evidence ofits involvement in early neural commitment, however,its role in the emergence and maturation of themammalian CNS remains unknown. Here, we inacti-vated Jmjd3 in the mouse and found that its losscauses perinatal lethality with the complete andselective disruption of the pre-Botzinger complex(PBC), the pacemaker of the respiratory rhythmgenerator. Through genetic and electrophysiologicalapproaches, we show that the enzymatic activity ofJMJD3 is selectively required for the maintenanceof the PBC and controls critical regulators of PBCactivity, uncovering an unanticipated role of this en-zyme in the late structuring and function of neuronalnetworks.

INTRODUCTION

Posttranslational modifications of histone tails have been asso-

ciated, and at times causally linked, to transcriptional outcomes.

Due to its comparative stability over other modifications (Byvoet

et al., 1972), lysine methylation was initially proposed as an ideal

candidate to mediate the epigenetic maintenance of tran-

scriptional states in both replicating and postmitotic cells. In

particular, di- and trimethylation of lysine 27 on histone H3

(H3K27me2/3) by EZH2 within Polycomb repressive complex 2

(PRC2) has emerged as a critical mechanism of gene repression

and, more recently, as a bona fide epigenetic mark that can be

propagated through cell division (Hansen et al., 2008;Margueron

et al., 2009). Two lines of experimentation revealed, however,

the unexpected dynamics of this mark: (1) the identification of

H3K27me2/3-specific JmjC-domain demethylases JMJD3 and

UTX (Agger et al., 2007; Burgold et al., 2008; De Santa et al.,

2007; Jepsen et al., 2007; Lan et al., 2007; Xiang et al., 2007);

and (2) the description of genome-wide changes in H3K27me3

that accompany the entire process of neural fate acquisition

(Mohn et al., 2008). The developing CNS has indeed become

a paradigm-setting model to interrogate how H3K27me3-

mediated repression times the activity of critical regulators of

neurogenesis, as shown through the ablation of Ezh2 at different

stages of corticogenesiswhosedivergent phenotypes of delayed

or accelerated neurogenesis (Hirabayashi et al., 2009; Pereira

et al., 2010) were recently reconciled in a model of Polycomb-

mediated developmental timing that reflects both the epigenetic

transmission of the mark and its dynamic relocation to new

targets (Testa, 2011). Conversely, the exquisite sensitivity of

neurogenesis to modulation of H3K27me3 underscores the rele-

vanceof timedH3K27me3demethylation, as revealed by the crit-

ical role of JMJD3 in both mouse embryonic stem cell (ESC)

neurulation (Burgold et al., 2008) and chick spinal cord neurogen-

esis (Akizu et al., 2010). Yet, despite these early advances the

relevanceof JMJD3 for the functionof themammalianCNS invivo

remains to be explored and nothing is known about its role in

later stages of CNS development and in neuronal maturation.

Cell Reports 2, 1–15, October 25, 2012 ª2012 The Authors 1

Please cite this article in press as: Burgold et al., The H3K27 Demethylase JMJD3 Is Required for Maintenance of the Embryonic Respiratory NeuronalNetwork, Neonatal Breathing, and Survival, Cell Reports (2012), http://dx.doi.org/10.1016/j.celrep.2012.09.013

Here, we inactivated Jmjd3 in the mouse to study its function

in CNS emergence and maturation, and we describe a specific

function of its enzymatic activity in the maintenance of the respi-

ratory rhythm generator (RRG).

The perinatal RRG is hypothesized to be formed by two

coupled oscillators, the retrotrapezoid nucleus/parafacial respi-

ratory group (RTN/pFRG) and the pre-Botzinger complex (PBC).

The PBC, ventral to the nucleus ambiguus (nA), functions from

embryonic day 15.5 (E15.5) and possibly constitutes the primary

respiratory oscillator (Fortin and Thoby-Brisson, 2009; Onimaru

et al., 2009; Smith et al., 1991; Thoby-Brisson et al., 2009). The

RTN/pFRG, ventral to the facial motor nucleus (nVII), functions

as early as E14.5, is CO2/pH chemosensitive and may be a

secondary oscillator (Thoby-Brisson et al., 2009).

Genetic manipulation led to the identification of several genes

that contribute to RRG emergence and function. MafB or Math1

inactivation drastically alters the PBC rhythm without abolishing

it (Blanchi et al., 2003; Rose et al., 2009). Inactivation of Phox2b,

Tshz3, or Task2 affects instead RTN/pFRG chemosensibility,

again disturbing but not abolishing PBC rhythm (Amiel et al.,

2009; Caubit et al., 2010; Gestreau et al., 2010). Inactivation of

Maoa, Phox2a, and Necdin affects the monoaminergic system,

which disturbs RRG maturation but does not silence the PBC

(Bou-Flores et al., 2000; Viemari et al., 2004; Zanella et al.,

2008). Finally, only two mutants were identified thus far with

totally silenced RRG, the VGlut2 mutant and the Dbx1 mutant,

where PBC neurons fail to achieve synchronous activation and

to elaborate central respiratory drive (Bouvier et al., 2010; Wal-

len-Mackenzie et al., 2006). Here, we show that JMJD3 enzy-

matic activity is critical for RRG maintenance and that JMJD3

controls the expression of specific markers and critical regula-

tors of PBC function. Lack of JMJD3 did not impair the early

emergence of the RRG at E15.5, but specifically altered the

maintenance of its PBC component, totally silencing it at E18.5

and leading to neonatal death with full penetrance.

RESULTS

Loss of JMJD3 Results in Perinatal LethalityWe inactivated Jmjd3 in the mouse starting from a gene trap

ESC line in which the first noncoding exon of the Jmjd3 transcript

had been trapped by a splice-acceptor (sA)-based knockout

cassette. We mapped its integration to intron 1 (Figure 1A) and

devised a PCR-based strategy to discriminate the heterozygous

from the homozygous state of the Jmjd3 knockout allele in the

mice that we derived through blastocyst injection (Figure 1B).

In heterozygous crosses, we recovered no homozygous knock-

out animals at weaning age (Figure 1C). We thus went back

to earlier developmental stages and retrieved homozygous

knockout fetuses up to embryonic day E18.5, indicating that

loss of JMJD3 results in perinatal or neonatal lethality. In the light

of JMJD3 function in early development, this surprising finding

prompted us to evaluate the impact of the gene trap cassette

on JMJD3 expression at the onset of development. We derived

ESC lines from heterozygous crosses and found that Jmjd3�/�

ESC still expressed nearly half the amount of JMJD3 as their

wild-type controls, suggesting that in ESC (Figure 1D) the trap

cassette had been partially circumvented through alternative

2 Cell Reports 2, 1–15, October 25, 2012 ª2012 The Authors

promoters. RT-PCR confirmed that while transcripts spanning

the exon1-exon2 junction were absent in Jmjd3�/� ESC, the

downstream exon combinations were still present (Figure 1E),

though Jmjd3 transcript levels were reduced by 4- to 5-fold

and decreased further to less than 10-fold during CNS develop-

ment (Figure 1F). We confirmed the virtually complete ablation of

Jmjd3 expression during development also by RNA in situ

hybridization with a probe located at the 30 end of the gene. As

shown in Figure 1G, at embryonic day E16.5 Jmjd3 is broadly

expressed, with particularly high levels in various sites of the

CNS, including the cortex, the basal ganglia, the olfactory epi-

thelium, and the cerebellar primordium. In mutant embryos,

Jmjd3 expression was instead almost completely ablated, indi-

cating that themost upstream promoter of Jmjd3 is the dominant

one in most tissues.

Jmjd3-Null Mice Die at Birth of Respiratory FailureIn order to determine the cause of perinatal lethality in Jmjd3-null

pups, we followed the natural delivery of several litters (n = 14)

taking care not to disturb maternal care. Jmjd3-null neonates

had normal weight and morphology but most of them did not

breathe, quickly became cyanotic, and died right after birth or

were found already dead (Figures 2A and 2B). A few survived

for several hours (n = 8/32, %12 hr) and only a tiny minority

survived longer than 12 hr but less than 1 day (n = 3/32, %

24 hr) (Figure 2A). In these few pups, the tissue architecture of

the lungs was unaffected, with normally inflated alveolar spaces,

thus excluding a primary lung defect as the cause of respiratory

failure (Figure S1). Exteriorization at E18.5 revealed that most

wild-type and heterozygous embryos breathed and survived,

whereas all Jmjd3-null embryos rapidly died, without producing

any discernible respiratory efforts. We further examined breath-

ing behavior in four litters from heterozygous crosses using

whole-body plethysmography (Figures 2C–2E). Following exteri-

orization at E18.5, Jmjd3+/+ and Jmjd3+/� embryos (n = 12/22 in

total) first gasped for 2–5min, producing deep respiratory move-

ments involving the whole-body muscles at a slow frequency

(about 2–5 cycles per minute [c.min]�1), and thereafter displayed

normal, shallow breathing at a faster frequency (46 ± 7 c.min�1;

Figure 2E). Plethysmography never revealed any gasps or

normal respiratory movements in Jmjd3�/� embryos (n = 8/8;

Figure 2E). To exclude heart failure as the primary cause for

death in Jmjd3�/� mice, we recorded the electrocardiogram

(ECG) of E18.5 embryos kept in utero (Figure S2). The heart

was regularly beating in Jmjd3�/� (n = 5) and Jmjd3+/+ or

Jmjd3+/� (n = 10) embryos (102 ± 17 c.min�1 and 68 ± 9

c.min�1, respectively). To exclude muscular or neuromuscular

dysfunctions as causes for the lack of breathing in Jmjd3-null

embryos, single electrical shocks were applied to the diaphragm

or the phrenic nerve of embryos and induced chest movements

(data not shown).

TheRespiratory RhythmGenerator DoesNot Function inJmjd3-Null Embryos at E18.5Given the proficiency of the cardiac and the peripheral respira-

tory systems in Jmjd3-null embryos, we suspected a deficit in

the RRG and used en blocmedullary preparations of exteriorized

embryos at E18.5 to examine whether the isolated RRG of

Figure 1. Inactivation of Jmjd3

(A) Scheme of the mouse Jmjd3 locus showing exons along with the integration site of the trap cassette, which comprises a splice acceptor (SA), the

b-galactosidase neomycin resistance gene fusion (b-geo) and a polyadenylation signal (pA). The probe used to assess Jmjd3 expression by RNA in situ

hybridization (Probein situ) is indicated by black bars below 30 terminal exons. The JmjC domain is shown mapped onto the relevant exons (17–20).

(B) Triplex PCR-based genotyping of wild-type (+/+), heterozygous (+/�), and homozygous (�/�) littermates (left). The scheme on the right shows the position of

the three primers, with the Forw1/Rev1 product amplified from the wild-type allele (282 bp), and the Forw1/Rev2 product amplified from the trapped allele

(259 bp).

(C) Table showing the number of wild-type, heterozygous and homozygous embryos at the indicated stage of development (E10.5–E18.5) and of pups at weaning

age (P21). Statistical analysis was performed with Chi-square test.

(D) Western blot analysis of JMJD3 expression in wild-type and homozygous mutant ESC clones. Vinculin served as loading control.

(E) Identification of Jmjd3 transcript variants by RT-PCR. In Jmjd3�/� ESC, no product was detected for the primer pair spanning exons 1 and 2. All other primer

combinations yielded a product in Jmjd3 mutant ESCs. Tbp served as housekeeping control.

(F) qRT-PCR analysis of Jmjd3 mRNA expression in ESCs, E13.5 brains and E16.5 forebrains and hindbrains. Tbp served as housekeeping control.

(G) RNA in situ hybridization showing Jmjd3 expression on sagittal sections of Jmjd3+/+ and Jmjd3�/� embryos at E16.5 of embryogenesis. Jmjd3+/+ embryos

(upper row) showed Jmjd3 expression in the cortex (cx), cerebellum (cb), mesencephalon (mes), basal ganglia (bag), and the olfactory epithelium (oe) as well as in

liver (li), gut (gut), and kidney (ki). In Jmjd3�/� mutants (lower row) the signal was absent or barely detectable, indicating virtually complete loss of the Jmjd3

transcript.

Please cite this article in press as: Burgold et al., The H3K27 Demethylase JMJD3 Is Required for Maintenance of the Embryonic Respiratory NeuronalNetwork, Neonatal Breathing, and Survival, Cell Reports (2012), http://dx.doi.org/10.1016/j.celrep.2012.09.013

Jmjd3-null embryos was able to produce respiratory-like phrenic

bursts (PBs) in vitro (Blanchi et al., 2003; Caubit et al., 2010) (Fig-

ure 3A). Rhythmic PBs were commonly recorded in all Jmjd3+/+

(n = 6/6) and most Jmjd3+/� (n = 14/18) preparations (12 ± 4

c.min�1 and 10 ± 1 c.min�1, respectively) but never in Jmjd3�/�

preparations (n = 8/8; Figures 3B and 3C).

Cell Reports 2, 1–15, October 25, 2012 ª2012 The Authors 3

Figure 2. Perinatal Lethality and Respiratory Failure upon Loss of JMJD3

(A) Table showing the number of Jmjd3+/+, Jmjd3+/�, and Jmjd3�/� neonates at birth (P0). Out of 32 Jmjd3�/� neonates, most (21) were born dead or died at birth,

few mice (eight) died within the first 12 hr postpartum and only three survived longer than 12 hr but less than 24 hr.

(B) Picture showing the appearance of neonates from heterozygous crosses immediately after birth. The two cyanotic neonates (first two in the row) were born

dead and afterward confirmed as Jmjd3�/�, a third Jmjd3�/� neonate survived for 20 hr (last pup in the row).

(C) Plethysmographic recordings of breathing activity in vivo from surgically delivered Jmjd3+/+ and Jmjd3�/� E18.5 fetuses. All wild-type mice initiated respi-

ratory cycles of inspirations (upward deflections) and expirations (downward deflections), whereas none of the Jmjd3�/� littermates showed any sign of

ventilation.

(D) Histogram showing for wild-type, heterozygous and mutant genotypes the percentage of surgically delivered E18.5 embryos that did (white bar) or did not

(black bar) show respiratory activity. The number of mice analyzed is indicated in each bar.

(E) Quantification of respiratory frequency indicated in respirations per minute. Bars represent the mean ± SEM of the numbers (n) of mice indicated on the x axis

below the graph.

See also Figure S1.

Please cite this article in press as: Burgold et al., The H3K27 Demethylase JMJD3 Is Required for Maintenance of the Embryonic Respiratory NeuronalNetwork, Neonatal Breathing, and Survival, Cell Reports (2012), http://dx.doi.org/10.1016/j.celrep.2012.09.013

The lack of rhythmic PBs in Jmjd3�/� preparations could result

from either an inactive RRG or a failure in synaptic transmission

from the RRG to the spinal phrenicmotor neurons. To distinguish

between these possibilities, we electrically stimulated either the

spinal cord or themedulla in Jmjd3�/� preparations. As depicted

in Figure 3D, we applied single electrical shocks at the level of

the second cervical segment of the spinal cord to activate the

respiratory output pathways running from the medullary res-

piratory areas toward the phrenic motor neurons (St2) and

recorded the activation of phrenic motor neurons at the level

of the fourth cervical segment (C4). This stimulation induced

a short latency response of phrenic motor neurons in both

Jmjd3�/� and Jmjd3+/+ preparations, revealing that Jmjd3�/�

phrenic motor neurons responded to spinal synaptic excitation

(Figure 3G). Second, we applied electrical shocks to the medulla

(St1 in Figure 3D), close to respiratory related areas (150–200 mm

below the surface of the ventrolateral medulla). In wild-type

preparations, as previously reported (Blanchi et al., 2003), med-

ullary stimulation induced brief responses of phrenic motor

neurons and, when delivered during the second third of expira-

tion, it shortened the silent expiratory period and precipitated

the onset of the next PB (Figure 3D, upper trace). Instead in

Jmjd3�/� preparations, the medullary stimulation never induced

a PB (Figure 3D, lower trace) but commonly induced brief

responses of phrenic motor neurons (Figure 3E). The phrenic

responses to spinal and medullary stimulations were observed

4 Cell Reports 2, 1–15, October 25, 2012 ª2012 The Authors

from both ipsilateral and controlateral sides, revealing normally

crossed pathways in Jmjd3�/� preparations. When applying

long train of shocks (3 s at 100 Hz) to either the ventrolateral

medulla or the median raphe area, we observed sustained and

long-lasting discharges of phrenic roots, persisting over several

seconds (Figure 3F), but never rhythmic PB.

Thus, given the ability of Jmjd3�/� phrenic motor neurons to

respond to synaptic inputs from spinal and medullary regions

and even to fire for long periods, the lack of rhythmic PB was

indicative of a defective or quiescent RRG. We attempted to

activate the RRG of Jmjd3�/� preparations by applying acidified

artificial cerebrospinal fluid (aCSF) (n = 2) or aCSF containing

either serotonin (25 mM; n = 2) or norepinephrine (25 mM; n = 2)

or elevated concentration of potassium (9 mM; n = 2) to the

medulla. None of these experimental maneuvers induced rhyth-

mic PB (data not shown), arguing for a defective rather than

quiescent RRG.

Disruption of the Pre-Botzinger Complex in E18.5Jmjd3-Null EmbryosWe next investigated whether the functional deficit in the

RRG could be traced to a neuroanatomical defect in Jmjd3�/�

embryos. For this, we conducted an immunohistochemical

analysis of the neurons organized in the two hypothesized

coupled oscillators of the murine RRG, the PHOX2b-positive

neurons of the RTN/pFRG and the Neurokinin 1 receptor

Figure 3. Lack of Respiratory Bursts in Jmjd3�/� Medullary In Vitro Preparations

Medullary in vitro (en bloc) preparations from Jmjd3�/� embryos surgically delivered at E18.5 do not produce rhythmic phrenic bursts although phrenic motor

neurons are synaptically excitable by electrical stimulations of the medulla and spinal cord.

(A) Paired traces show raw (C4) and integrated (IntC4) signals recorded from the phrenic nerve of medullary in vitro preparations isolated from Jmjd3+/+ and

Jmjd3�/� embryos (upper and lower traces, respectively) surgically delivered at E18.5.

(B) Histogram showing the percentage of Jmjd3+/+, Jmjd3+/�, and Jmjd3�/� E18.5 brainstem preparations that did (white bars) or did not (black bars) produce

rhythmic phrenic bursts (the number of preparations analyzed is indicated in each bar).

(C) Quantification of phrenic burst frequency expressed in cycles per minute. Bars represent the mean ± SEM of the numbers (n) of mice indicated on the x axis

below the graph.

(D) Schematic drawing of the en bloc embryonic preparation (ventral surface upward) showing the medullary (St1) and spinal (St2) stimulation sites and the

phrenic nerve recording site at the level of the C4 ventral root (C4). Upper and lower traces show the presence and absence of rhythmic phrenic bursts (integrated

C4 signal) in preparations of Jmjd3+/+ and Jmjd3� E18.5 embryos, respectively. Applying a single shock electrical stimulation (1 ms, 2 V; black triangle) at the

medullary St1 site during expiration (silent phrenic interval) triggered a premature phrenic burst in Jmjd3+/+ but not in Jmjd3�/� preparations. Inserts (enlarged

time scale: 1 s) show spontaneous and stimulation-induced phrenic bursts in the Jmjd3+/+ preparation and the lack of phrenic burst after stimulation in the

Jmjd3�/� preparation.

(E) In E18.5 Jmjd3�/� preparations, superimposed traces (raw C4 signal) show that single shock stimulations delivered in the medulla at St1 (1 ms, 2 V; black

triangle) induced brief synaptic responses of phrenic motor neurons (arrows).

(F) In E18.5 Jmjd3�/� preparations, applying repetitive stimulation (100 Hz for 3 s; black bar) to St1 induced long-lasting discharges (integrated C4 signal) of the

previously silent phrenic motor neurons, but no rhythmic phrenic bursts. The amplitude and the duration of the stimulation-induced discharges depended on the

stimulus strength (from top to bottom, 3, 1.5 and 1 V; pulse duration 1 ms). Dotted lines indicate baseline level of the C4 signal prior to stimulation.

(G) as in (E), but following stimulation of the spinal cord at the level of St2 (1 ms, 2 V; black triangle). In E18.5 Jmjd3�/� preparations, superimposed traces (raw C4

signal) show that single shock stimulations at St2 induced short-lasting responses of phrenic motor neurons (arrows).

See also Figure S2.

Please cite this article in press as: Burgold et al., The H3K27 Demethylase JMJD3 Is Required for Maintenance of the Embryonic Respiratory NeuronalNetwork, Neonatal Breathing, and Survival, Cell Reports (2012), http://dx.doi.org/10.1016/j.celrep.2012.09.013

(NK1R)/Somatostatin (SST)-positive neurons of the PBC. The

embryonic PBC has been previously defined through NK1R

and SST expression below the nA, extending toward the ventral

surface of the medulla at the level of the inferior olive (Blanchi

et al., 2003; Caubit et al., 2010; Gray et al., 2001, 2010; Stornetta

et al., 2003). We confirmed this PBC-typical distribution in wild-

type embryos but not in Jmjd3�/� embryos where the corre-

sponding area appeared grossly altered in extension and struc-

ture. As shown in Figure 4A, NK1R-positive (left column) and

SST-positive (middle column) neurons were present ventrally

of the nA of Jmjd3�/� medulla, but they were significantly fewer

than in the wild-type counterpart, with a much weaker density

Cell Reports 2, 1–15, October 25, 2012 ª2012 The Authors 5

Figure 4. Neuroanatomical Characterization of the Pre-Botzinger Complex in Jmjd3�/� Embryos

(A) Immunohistochemical staining for NK1R (first column) and SST (second column) on coronal sections from wild-type (upper row) and Jmjd3�/� (lower row)

E18.5 brainstems. A merged image of NK1R (magenta) and SST (green) stainings is shown in the last column on the right.

(B) Three-dimensional reconstruction (in sagittal view, left panels) of the NK1R-positive area comprising the nA and the PBC on consecutive coronal sections of

Jmjd3+/+ and Jmjd3�/� E18.5 brainstems. Volumetric measurement, showing the mean ± SD from five wild-type and four mutant E18.5 samples, is shown in the

right panel (**p < 0.01, Student’s t test).

(C) Immunohistochemical staining for SST (first column) and PAX2 (second column) on coronal sections from wild-type (upper row) and Jmjd3�/� (lower row)

E18.5 brainstems. A merged image of PAX2 (magenta) and SST (green) is shown in the last column on the right.

(D) Graphs showing the number of SST- (left graph) and PAX2- (right graph) positive cells per coronal hemisection at various levels (along the rostrocaudal axis,

x axis) of the rostral ventrolateral Jmjd3+/+ and Jmjd3�/� E18.5 medulla. Cells were counted within an area of 2603 290 mm encompassing the PBC core region

6 Cell Reports 2, 1–15, October 25, 2012 ª2012 The Authors

Please cite this article in press as: Burgold et al., The H3K27 Demethylase JMJD3 Is Required for Maintenance of the Embryonic Respiratory NeuronalNetwork, Neonatal Breathing, and Survival, Cell Reports (2012), http://dx.doi.org/10.1016/j.celrep.2012.09.013

Please cite this article in press as: Burgold et al., The H3K27 Demethylase JMJD3 Is Required for Maintenance of the Embryonic Respiratory NeuronalNetwork, Neonatal Breathing, and Survival, Cell Reports (2012), http://dx.doi.org/10.1016/j.celrep.2012.09.013

and a diffuse, unstructured distribution of the NK1R and

SST signals. This was particularly prominent for the subset of

NK1R/SST coexpressing neurons, which was recently defined

as the core of the PBC in mice (Gray et al., 2010). Complete

acquisition of coronal sections enabled the three-dimensional

reconstruction and volumetric measurement of the NK1R-posi-

tive network of neurons comprising the nA and the PBC (Fig-

ure 4B; Movie S1), confirming the stark alterations in the volume

and shape of the PBC in Jmjd3�/� samples. These abnormalities

were reflected also in the distribution of PBC neurons coex-

pressing SST with PAX2, one of three transcription factors

(along with PHOX2b and LHX9) expressed in the ventrolateral

medulla (VLM) in specific neuronal subsets (Gray et al., 2004,

2010). As shown in Figure 4C, the number of SST/PAX2 coex-

pressing neurons was severely reduced in Jmjd3�/� PBCs, as

determined by serial sectioning and counting along the rostro-

caudal axis of the medulla. Given that, in the same regions of

Jmjd3�/� ventrolateral medulla, the overall number of PAX2-

positive neurons was only marginally affected (with a significant

reduction only in the uppermost caudal PBC, Figure 4D), the

selective alteration in SST/PAX2-double-positive neurons points

to a specific requirement for JMJD3 in this specific neuronal

subpopulation. Consistently, we also did not observe any signif-

icant decrease in the overall number of glutamatergic neurons,

as identified through VGLUT2 expression (Figure S3A). Finally,

global levels of H3K27me3 were not affected in mutant PBC

neurons, as expected on the basis of convergent evidence

from several cell types in which JMJD3 was shown to control

locus-specific rather than global H3K27me3 (De Santa et al.,

2009) (Figure S3B).

In the light of the neuroanatomical alterations of the PBC,

we asked whether loss of JMJD3 affected also the RTN/pFRG.

As shown in Figure S3C, we identified the expected group

of PHOX2b-expressing neurons in the RTN/pFRG area, ven-

trally of the VII motor nucleus, in both Jmjd3+/+ and Jmjd3�/�

embryos, and no differences were observed in PHOX2b and

NK1R expression and in the overall structure of the RTN/pFRG

(Figure S3C). Consistently, we found the expression of Jmjd3

in the brainstem to be predominant in the PBC, nA, and nVII

and comparatively much lower if barely detectable in the RTN/

pFRG (Figures 4E and S3D). Thus, the differential involvement

of the twomedullary oscillators in ourmutants points to a specific

requirement for JMJD3 in PBC integrity at E18.5, consistent with

its predominant expression pattern.

(six PBC/genotype). Shown is, for the value of each level, the mean ± SD; asteris

t test).

(E) Overlay of immunohistochemical staining for NK1R (red) and Jmjd3 (RNA in

performed on consecutive sagittal sections fromwild-type E18.5 brainstem and th

medulla.

(F) Histogram showing the percentage of Jmjd3+/+, Jmjd3+/� and Jmjd3�/� pre

produce rhythmic diaphragmatic EMG bursts at E15.5, E16.5 and E18.5. The num

preparations with rhythmic EMG bursts decreased with maturational age in Jmjd3

(G) Immunohistochemical staining for NK1R (first column) and SST (second colu

(H) Histogram showing the brainstem content of the 5-HT precursor L-Trp, 5-HT an

8) embryos at E16.5. Brainstem levels of 5-HT and 5-HIAA were significantly lowe

content was similar for all three genotypes. Bars represent the mean ± SEM and a

Scale bars represent 100 mm (A, C, G) and 10 mm (E). nA: nucleus ambiguus, PB

See also Figure S3.

JMJD3 Is Required for the Maintenance of thePre-Botzinger ComplexThe embryonic RRG emerges functionally at embryonic days

E15.5 and E14.5, respectively, for the coupled PBC and RTN/

pFRG respiratory oscillators (Thoby-Brisson et al., 2005, 2009).

Because our immunohistochemical analysis showed an altered

PBC in the presence of a normal RTN/pFRG at day E18.5, we

conducted electrophysiological experiments at earlier embry-

onic stages to probe the role of JMJD3 in the functional emer-

gence of both oscillators.

At E16.5, a total of 70 embryos from six litters were analyzed

using in vitro medullary preparations retaining the rib cage to

allow visualization of respiratory movements and recording of

chest electromyogram (EMG) discharges (Ren et al., 2003).

Rhythmic chest movements were observed in most wild-type

and Jmjd3+/� preparations (n = 14/16 and 32/34, respectively)

as well as in half of the Jmjd3�/� preparations (n = 10/20; Fig-

ure 4F, middle graph), demonstrating that the embryonic RRG

can function at this early embryonic age despite the lack of

JMJD3. We then extended our study to day E15.5, when most

preparations produced rhythmic chest movements, indepen-

dent of genotype (n = 6/7, 12/13 and 3/4 for Jmjd3+/+, Jmjd3+/�,and Jmjd3�/� preparations, respectively; Figure 4F, left graph).

Together, these results show that whereaswith embryonicmatu-

ration the ratio of active/inactive preparations remained constant

in Jmjd3+/+ and Jmjd3+/� embryos (about 80%, n = 7/8 and 14/

18, respectively), it significantly decreased in Jmjd3�/� embryos,

dropping from 75%at E15.5 to 50%at E16.5 to 0%at E18.5 (Fig-

ure 4F, right graph). Figure S3E shows representative EMG re-

cordings of Jmjd3+/+ and Jmjd3�/� preparations at day E15.5,

and the complete lack of activity in day E18.5 Jmjd3�/� prepara-

tions. In order to further dissect the role of JMJD3 in the emer-

gence of the RRG, we then probed the functional features of

the chemosensing neurons that were reported to reside in the

maturing RTN/pFRG and to be responsive to pH changes from

as early as day E15.5–E16.5 (Caubit et al., 2010). We applied

acidified aCSF to the medulla of four Jmjd3+/+ and two Jmd3�/�

preparations and observed that acidosis increased the fre-

quency of respiratory EMG discharges in half the samples of

either genotype (n = 2/4 and n = 1/2 for Jmjd3+/+ and Jmd3�/�

preparations, respectively; data not shown).

Finally, we examined the neuroanatomical correlate of the

apparently normal functional emergence of the RRG, analyzing

the structure of the PBC in Jmjd3�/� embryos at day E16.5.

ks indicate statistically significant differences (*p < 0.05, **p < 0.01, Student’s

situ hybridization, blue). NK1R staining and Jmjd3 in situ hybridization were

e images from the same area were alignedwith respect to the ventral limit of the

parations with attached rib cage that did (white bars) or did not (black bars)

ber of preparations analyzed is indicated in each bar. The percentage of active�/� embryos, whereas it remained constant in Jmjd3+/+ and Jmjd3+/� embryos.

mn) on coronal sections from Jmjd3+/+ and Jmjd3�/� E16.5 brainstems.

d themetabolite 5-HIAA in Jmjd3+/+ (n = 8), Jmjd3+/� (n = 10) and Jmjd3�/� (n =

r in Jmjd3�/� embryos than in Jmjd3+/+ and Jmjd3+/� embryos, whereas L-Trp

sterisks indicate statistically significant differences (p < 0.05, Student’s t test).

C: pre-Botzinger complex, VII: facial motor nucleus.

Cell Reports 2, 1–15, October 25, 2012 ª2012 The Authors 7

Figure 5. Exogenous Expression of Jmjd3 Rescues the Perinatal Lethality of Jmjd3�/� Mutants

(A) Schematic illustration of the targeting strategy to insert the Jmjd3 transgene into the Rosa26 locus. The targeting vector contains a splice acceptor site (SA),

a loxP-flanked STOP cassette, the Jmjd3 coding sequence followed by an FRT-flanked IRES-EGFP cassette and a polyadenylation sequence (pA), resulting in

Cre-mediated expression of Jmjd3 from the endogenous Rosa26 promoter.

(B) Table showing the number of Jmjd3+/+, Jmjd3+/�, and Jmjd3�/� mice carrying the Rosa26Jmjd3 allele that were recovered at weaning age (P21).

(C) Immunohistochemical staining for NK1R (first column) and SST (second column) on coronal sections from Jmjd3�/� (upper row) and Jmjd3�/�;Rosa26Jmjd3

(lower row) E18.5 brainstems. A merged image of NK1R (magenta) and SST (green) is shown in the last column on the right. Scale bars represent 100 mm.

See also Figure S4.

Please cite this article in press as: Burgold et al., The H3K27 Demethylase JMJD3 Is Required for Maintenance of the Embryonic Respiratory NeuronalNetwork, Neonatal Breathing, and Survival, Cell Reports (2012), http://dx.doi.org/10.1016/j.celrep.2012.09.013

Staining for NK1R revealed a comparatively better preservation

of the emerging PBC structure in homozygous null embryos

with respect to the striking defects observed at day E18.5,

consistent with the largely normal function of the emerging

PBC in the absence of JMJD3 (Figure 4G). Interestingly, the

SST pattern was more severely affected already at E16.5, sug-

gesting either an earlier sensitivity of the NK1R/SST-double-

positive neuronal subset to the loss of JMJD3, or a differential

impact of JMJD3 on the expression of the two genes. Finally,

we did not find evidence of caspase-dependent apoptosis in

mutant PBC (data not shown), in agreement with the broadly

conserved number of PAX2-positive neurons recovered at

E18.5. Together with the distribution of VGLUT2-positive neu-

rons, this confirms that loss of JMJD3 did not cause widespread

neuronal loss in the ventral medulla but impacted selectively on

the integrity of the NK1R/SST neuronal core within the PBC.

The emergence, function, and maturation of the PBC starting

from day E15.5 has been associated with the input of serotonin

(5-HT)-specific neurons from the developing midline raphe sys-

tem, consistent with the well-established role of early 5-HT sig-

nal in networks maturation (Gaspar et al., 2003), including RRG

maturation (Bou-Flores et al., 2000; Hilaire et al., 2010) and initi-

ation of rhythmic activity in the developing hindbrain (Hunt et al.,

2005). We thus investigated the levels of tryptophan, serotonin

8 Cell Reports 2, 1–15, October 25, 2012 ª2012 The Authors

and its catabolyte, 5-hydroxyindolacetic acid (5-HIAA), in the

brainstem of E16.5 Jmjd3 mutant and control embryos and

found a significant decrease of both 5-HT and 5-HIAA upon

loss of JMJD3, suggesting that loss of JMJD3 may impair

PBC maturation also through defective formation of the 5-HT

system (Figure 4H). However, since the development and func-

tion of the RTN/pFRG was not affected by JMJD3 loss, we must

conclude that PBC failure results mainly from the selective

requirement of JMJD3 for its maintenance.

Exogenous Expression of JMJD3 Rescues the PerinatalLethality Caused by the Jmjd3 Knockout AlleleIn order to gain unequivocal evidence that the respiratory failure

and the lack of PBCmaintenance in our mutants were caused by

lack of JMJD3, we sought to rescue the phenotype by re-ex-

pressing Jmjd3. To this end, we engineered a mouse strain for

the Cre-loxP-dependent expression of Jmjd3 from the Rosa26

(R26) genomic locus (Figure 5A). Following identification of

homologous recombinant ESC clones (Figure S4), we derived

chimeras that transmitted the allele through the germline (here-

after Rosa26flSTOP-Jmjd3). We then bred the Rosa26flSTOP-Jmjd3

allele into the Jmjd3 knockout strain along with a ubiquitous

Cre-deleter strain to check whether the recombined Jmjd3

transgene (Rosa26Jmjd3) was able to rescue perinatal lethality.

Please cite this article in press as: Burgold et al., The H3K27 Demethylase JMJD3 Is Required for Maintenance of the Embryonic Respiratory NeuronalNetwork, Neonatal Breathing, and Survival, Cell Reports (2012), http://dx.doi.org/10.1016/j.celrep.2012.09.013

As shown in Figure 5B, mice homozygous for the gene trap allele

(Jmjd3�/�) and harboring the Rosa26Jmjd3 allele reached wean-

ing age at the correct Mendelian ratios. While this provided

conclusive evidence that the expression of JMJD3 suffices

to rescue RRG function, we confirmed this rescue also with

the immunohistochemical characterization of a normal PBC in

Jmjd3�/�;Rosa26Jmjd3 compound mutants at E18.5 (Figure 5C).

Maintenance of the Respiratory Network Requires theEnzymatic Activity of JMJD3The relative inefficiency of demethylation catalyzed by JmjC

domain-containing enzymes, along with their often limited con-

tribution to the global levels of histone methylation and gene

transcription, led us to hypothesize that JmjC domain-containing

demethylases may function also independent of enzymatic ac-

tivity (Natoli et al., 2009). Genome-wide analysis in macrophages

confirmed that JMJD3 could regulate gene expression also in

a demethylation-independent manner (De Santa et al., 2009),

as subsequently established also for its demethylase-indepen-

dent role in chromatin remodeling (Miller et al., 2010) and more

recently for its regulation of gene expression in promyelocytic

leukemia (Chen et al., 2012). We thus aimed at determining

whether the catalytic function of JMJD3 was required for the

maintenance of the PBC. To this end, we isolated a bacterial

artificial chromosome (BAC) harboring the Jmjd3 locus (Fig-

ure 6A) and engineered through ‘‘recombineering’’ a point muta-

tion resulting in a Histidine to Alanine substitution at codon 1388,

which abolishes catalysis by the JmjC domain. We screened

oocyte injection-derived transgenic mutant-Jmjd3 mice (here-

after Jmjd3mut) by PCR (Figure 6B) and estimated the number

of integrations by Southern blot (Figure 6C), using a probe that

enabled comparison of the signal from the endogenous se-

quence with that from the BAC-specific one (Figure 6A). As the

Southern probe was not located in close proximity of the

Jmjd3 gene, we confirmed the copy number range with a quanti-

tative PCR (qPCR) probe spanning the second exon of Jmjd3

and hence providing a direct estimate of the number of extra

Jmjd3 integrated copies. As shown in Figure 6D, qPCR con-

firmed that Jmjd3mut21 and Jmjd3mut65 had, respectively, four

and one extra copies of the Jmjd3 locus that were transmitted

stably across generations (Figure 6D).

Next, we confirmed the expression of the mutant Jmjd3 allele

by sequencing a RT-PCR product spanning the mutated codon

and amplified from the brain of Jmjd3mut21 and Jmjd3mut65

mice (Figure 6E). Finally, we intercrossed mice heterozygous

for the Jmjd3 knockout allele (Jmjd3+/�) and harboring either

Jmjd3mut21 or Jmjd3mut65. As shown in Figure 6F, despite the

expected segregation of either Jmjd3mut allele among both

Jmjd3+/+ and Jmjd3+/� animals, we were unable to retrieve at

weaning age any compound Jmjd3�/�;Jmjd3mut21 or Jmjd3�/�;Jmjd3mut65 mutants, whereas we could readily obtain them

at embryonic day E18.5 (Figure S5). Importantly, in the brains

of both these E18.5 Jmjd3�/�;Jmjd3mut21 and Jmjd3�/�;Jmjd3mut65 compound mutants, we confirmed that the BAC

transgenes rescued the expression of Jmjd3 at least to the level

observed in Jmjd3�/�;Rosa26Jmjd3 embryos (Figure 6G). This

allows us to conclude that the expression of catalytically inactive

JMJD3 cannot compensate for the loss of the wild-type protein.

JMJD3 Regulates Critical PBC-Specific Genes throughH3K27me3 DemethylationTo gain mechanistic insight into the downstream targets through

which the enzymatic activity of JMJD3 controls PBC mainte-

nance and function, we started from the recent identification,

by cDNA subtraction, of thus far the most extensive subset of

rodent PBC-specific markers. Besides identifying several PBC

markers in addition to NK1R and SST, this study characterized

the glycoprotein Reelin for its high and selective expression in

a subset of PBC NK1R- and SST-positive neurons and for its

requirement in PBC-mediated response to hypoxia (Solomon

et al., 2000; Tan et al., 2012).We thus carried out ameta-analysis

of available data sets to interrogate the dynamics of H3K27me3

at the promoter of these PBC-specific genes during neural

differentiation, reasoning that this would guide us in predicting

which PBC-relevant genes are most likely impacted by failed

H3K27me3 demethylation. Importantly, this meta-analysis in-

cluded expression and genome-wide chromatin profiles from

the homogeneous differentiation of ESC into terminally dif-

ferentiated glutamatergic neurons (N) through intermediate cell

aggregate (CA) and neural precursor (NP) stages, a relevant

experimental system given the glutamatergic nature of PBC

neurons and the one that first exposed H3K27me3 dynamics

during neuronal differentiation (Bibel et al., 2004; Mohn et al.,

2008). We retrieved several PBC-specific genes that lost

H3K27me3 in the transition from ESC to N, including Reelin

(Reln), membrane-associated Ring Finger 4 (March4), kin of

IRRE-like 3 (Kirrel3, also known asNeph2), and estrogen-related

receptor gamma (Esrrg) and set out to validate the role of JMJD3

in their regulation starting from wild-type and Jmjd3�/� ESC. As

shown in Figure 7A, all four geneswere virtually silent in wild-type

ESC, peaked in NP and returned to intermediate levels in N,

but their upregulation was virtually completely abolished upon

loss of JMJD3. We next examined the dynamics of H3K27me3

in ESC, CA, NP, and N (Figure 7B). While in wild-type cells

H3K27me3 (controlled for H3 occupancy) underwent a dramatic

decrease in the transition from ESC through CA to NP and N, its

levels dropped significantly less (Kirrel3 and Reln) or remained

stable (March4 and Esrrg) in Jmjd3�/� cells, indicating impaired

H3K27me3 demethylation (Figure 7B). Consistently, and in

agreement with its expression pattern (data not shown), JMJD3

recruitment peaked at the CA stage and was severely com-

promised in mutants (Figure 7C).

We can thus conclude that during the in vitro acquisition of

glutamatergic fate, a hallmark feature of PBC neurons required

for rhythm generation, JMJD3 controls the expression of critical

PBC-specific genes through H3K27me3 demethylation, a func-

tion that we hypothesize to become eventually rate-limiting for

the maintenance of PBC function in vivo. Reln and Kirrel3 are

in this respect particularly relevant targets, given the involvement

of the former in PBC-mediated hypoxic response and the role of

the latter in structuring the pontine nucleus by controlling the late

stages of neuronal migration (Nishida et al., 2011).

DISCUSSION

The dynamics of Polycomb marking during neurogenesis sug-

gested a critical role for timed H3K27 demethylation in the

Cell Reports 2, 1–15, October 25, 2012 ª2012 The Authors 9

Figure 6. The Enzymatic Activity of JMJD3 Is Required for the Maintenance of the Respiratory Network

(A) Schematic representation of the BAC harboring the Jmjd3 locus engineered with a point mutation. Shown are the pBACe3.6 vector (gray oval), Jmjd3 exons

(white numbered rectangles), oligos used for genotyping (FGt and RGt), oligos used to amplify cDNA for sequencing of the point mutation (FRTseq and RRTseq),

probe used for Southern blot (ProbeSouthern), TaqMan copy number probe (ProbecnJmjd3), and TaqMan gene expression probe (ProbeExpression).

(B) PCR-based screening of Jmjd3mut transgenic mice, showing the amplification of a Jmjd3mut BAC-specific 710 bp PCR product in two independent founder

lines (Jmjd3mut21 and Jmjd3mut65).

10 Cell Reports 2, 1–15, October 25, 2012 ª2012 The Authors

Please cite this article in press as: Burgold et al., The H3K27 Demethylase JMJD3 Is Required for Maintenance of the Embryonic Respiratory NeuronalNetwork, Neonatal Breathing, and Survival, Cell Reports (2012), http://dx.doi.org/10.1016/j.celrep.2012.09.013

Please cite this article in press as: Burgold et al., The H3K27 Demethylase JMJD3 Is Required for Maintenance of the Embryonic Respiratory NeuronalNetwork, Neonatal Breathing, and Survival, Cell Reports (2012), http://dx.doi.org/10.1016/j.celrep.2012.09.013

acquisition of neural fate (Mohn et al., 2008; Testa, 2011). Direct

regulation by the H3K27 demethylase JMJD3 of key neurogenic

factors during mouse ESC neurulation and chick spinal cord

development confirmed the role of this enzyme in early neural

development (Akizu et al., 2010; Burgold et al., 2008); yet its

in vivo function in the emergence andmaturation of the mamma-

lian CNS remained unexplored. Through genetic inactivation of

Jmjd3 in the mouse, our work uncovers an unanticipated role

of this enzyme in the maintenance of the embryonic respiratory

neuronal network, extending the relevance of its enzymatic

activity from early lineage choices to the late structuring of

neuronal networks.

We found that truncation of the Jmjd3 transcript downstream

of its first noncoding exon results in perinatal lethality with full

penetrance. This finding, apparently at odds with JMJD3 re-

quirement for early developmental choices (Canovas et al.,

2012), suggests that the partial efficiency of the trap cassette

in ESC guarantees levels of the protein that apparently suffice

to carry through the earliest stages of development. Interest-

ingly, selective ablation of the JmjC domain also resulted in peri-

natal lethality, indicating that at least the enzymatic portion of

JMJD3 may be surprisingly dispensable for early mouse devel-

opment (Satoh et al., 2010).

Our findings establish that JMJD3 is selectively required for

the maintenance and function of the PBC component of the

RRG and uncover critical PBC-specific genes that are targets

of JMJD3 during the acquisition of glutamatergic neuronal

fate. Electrophysiological and immunohistochemical studies of

Jmjd3-null embryos at E15.5, E16.5, and E18.5 revealed that

the early stages of RRG development took place normally

despite the lack of JMJD3. In Jmjd3-null embryos at E15.5, the

RRG was able to produce a rhythmic respiratory drive and to

adapt it to pH changes, revealing a normal function of the

RTN/pFRG and PBC oscillators just after their emergence.

From E15.5 onward, however, lack of JMJD3 impaired progres-

sively the further maturation and rhythmogenicity of the PBC,

which became silent in half of Jmjd3-null embryos at E16.5 and

in all Jmjd3-null embryos at E18.5. This functional decay upon

loss of JMJD3 was mirrored in the network structure of the

PBC, which was still comparatively well preserved at E16.5 but

appeared severely disrupted at E18.5, with dramatic defects in

thenumber andarchitecture ofNK1R- andSST-positive neurons.

Finally, we identified several PBC-specific genes that depend

on JMJD3 for H3K27 demethylation-dependent upregulation

during neuronal differentiation. Importantly, these include Reelin,

(C) Southern blot of EcoRV-digested genomic DNA from Jmjd3mut21, Jmjd3mut65

Jmjd3mut BAC yielded fragments of, respectively, 7.7 and 5.1 kb.

(D) Quantitation of the number of BAC copies integrated in the Jmjd3mut21 and Jm

ProbecnJmjd3. The scheme illustrates the number of BAC-harbored Jmjd3 copies p

copies in total) and Jmjd3mut65 strain (three copies in total). Bars for (p) represen

(E) Sequencing trace of the RT-PCR product amplified from the brain, spanning ex

in gray) resulting in the Histidine to Alanine substitution at position 1388.

(F) Table showing the lack of phenotypic rescue by catalytically inert JMJD3,

heterozygous crossings between Jmjd3+/� mice carrying either the Jmjd3mut2

square test.

(G) qRT-PCR analysis of Jmjd3 mRNA expression in brains from E18.5 Jmjd

Jmjd3mut65 embryos.

See also Figure S5.

coding for the glycoprotein involved in PBC response to hypoxia,

and Kirrel3, a member of the immunoglobulin superfamily in-

volved in the late phases of neuronal migration underlying the

architecture of hindbrain nuclei. Thus, in light of the paucity of

functionally validated PBC markers, our approach also estab-

lishes the differentiation of ESC into glutamatergic neurons as

a paradigm to uncover and validate PBC-relevant genes.

Several features of this phenotype appear particularly striking

in relation to previously reported gene-specific respiratory defi-

cits. The first is the selectivity in PBC impairment. Thus, muta-

tions in the transcription factor DBX1 and the axon guidance

receptor ROBO3 produced no rhythmic phrenic bursts because

of alterations of brainstem crossed pathways impairing the pro-

duction of the respiratory central drive toward phrenic moto-

neurons (Bouvier et al., 2010). Crossed pathways were instead

spared in Jmjd3 mutants, just like the RTN/pFRG nucleus.

Second, the severity of the phenotype, with the complete lack

of RRG activity at E18.5, also stands in contrast with the pheno-

types observed in most other respiratory mutants, including

Mash1, Phox2a, Phox2b, Rnx, Hoxa1, Krox20, Ret, and MafB,

which presented in varying degrees severe aberrations but

not a complete absence of respiratory activity. This difference

appears particularly relevant in the case of the transcription

factor MAFB, whose phenotypic features (perinatal gasping

behavior coupled to an abnormally structured PBC) led to posit

the emergence in late gestation of rhythmogenic MAFB-depen-

dent neurons (Blanchi et al., 2003). However, despite the fact

that also JMJD3 is required for the late structuring of the PBC,

we can exclude a downstream involvement of MAFB in our

mutants since in MafB-null E18.5 embryos the RRG was active

and produced a gasping-like rhythm (Blanchi et al., 2003). The

observation that this along with most other respiratory mutants

maintained the generation of inspiratory bursts, albeit at much

reduced frequency, led to the suggestion that the mechanisms

underlying the production of such bursts may be particularly

resilient even in the context of aberrantly developed respiratory

circuits (Blanchi et al., 2003; Wallen-Mackenzie et al., 2006).

Indeed, the only mutant with a complete ablation of synchro-

nized PBC activity had been thus far that resulting from the

inactivation of vesicular glutamate transporter 2 (VGlut2) (Wal-

len-Mackenzie et al., 2006). Critically, however, in this mutant

the PBC never began functioning although it appeared mor-

phologically normal. Hence loss of JMJD3 represents the first

instance of a complete ablation of PBC activity which intervenes

however after its normal emergence at day E15.5.

and C57BL/6 (WT control) mouse tails. The endogenous Jmjd3 locus and the

jd3mut65 founders (f) and their progenies (p) assessed by real-time PCR using

resent in the wild-type control (two endogenous copies), Jmjd3mut21 strain (six

t the mean ± SD of n = 5 mice.

ons 17 and 18 and demonstrating expression of the point mutation (highlighted

assessed by genotyping mice recovered at weaning age and derived from

1 or the Jmjd3mut65 transgene. Statistical analysis was performed with Chi-

3+/+, Jmjd3�/�, Jmjd3�/�;Rosa26Jmjd3, Jmjd3�/�;Jmjd3mut21, and Jmjd3�/�;

Cell Reports 2, 1–15, October 25, 2012 ª2012 The Authors 11

Figure 7. JMJD3 Regulates PBC-Relevant Genes during the Acquisition of Glutamatergic Neuronal Fate

(A) Expression profiles of Esrrg,Kirrel3,March4, and Reln during differentiation of wild-type and Jmjd3�/� ESC into glutamatergic neurons through cell aggregate

and neural precursor intermediates. Shown are the means ± SD of qRT-PCR triplicates normalized to Tbp.

(B) Enrichment for H3K27me3 at the promoters of Esrrg, Kirrel3,March4, and Reln in ESC, CA, NP, and N along with H3 occupancy. Bars represent the means ±

SD of qPCR triplicates in a representative ChIP experiment.

(C) Assessment of JMJD3 binding to the promoters of Esrrg, Kirrel3,March4, and Reln in ESC, CA, NP, and N. ESC, embryonic stem cells;CA, cell aggregates at

day 8 of differentiation;NP, neural precursors following plating of CA at day 10 of differentiation;N, neurons. Bars represent themeans ± SD of qPCR triplicates in

a representative ChIP experiment.

Please cite this article in press as: Burgold et al., The H3K27 Demethylase JMJD3 Is Required for Maintenance of the Embryonic Respiratory NeuronalNetwork, Neonatal Breathing, and Survival, Cell Reports (2012), http://dx.doi.org/10.1016/j.celrep.2012.09.013

This last feature involves the role of JMJD3 in the mainte-

nance or late maturation of the PBC and appears particularly

relevant in the light of genetic evidence that defined for the

Trithorax family of proteins a selective role in the maintenance

rather than in the establishment of gene expression patterns

and body segment identities (Glaser et al., 2006; Kennison,

1995; Yu et al., 1998). Given our previous characterization of

the association between JMJD3 and the core components of

the Trithorax MLL1/4 complexes (De Santa et al., 2007), the

12 Cell Reports 2, 1–15, October 25, 2012 ª2012 The Authors

late requirement for JMJD3 in PBC structure and function

extends now the Trithorax paradigm from the maintenance of

gene expression patterns in early development to the stability

of neuronal networks in late gestation.

Finally, previous studies, including our own, had suggested

that JMJD3 could also function independent of its enzymatic

activity (Burgold et al., 2008; De Santa et al., 2009; Miller et al.,

2010; Natoli et al., 2009). Thus, it became critical to test whether

the role of JMJD3 in PBC maintenance requires its enzymatic

Please cite this article in press as: Burgold et al., The H3K27 Demethylase JMJD3 Is Required for Maintenance of the Embryonic Respiratory NeuronalNetwork, Neonatal Breathing, and Survival, Cell Reports (2012), http://dx.doi.org/10.1016/j.celrep.2012.09.013

activity. Our results provide conclusive evidence that it does,

uncovering a vital function for the catalytic properties of this

protein.

EXPERIMENTAL PROCEDURES

Murine Strains

Experiments involving animals were performed in accordance with the Italian

and French Laws that enforce EU 86/609 Directive (Council Directive 86/609/

EEC of 24 November 1986 on the approximation of laws, regulations and

administrative provisions of the Member States regarding the protection of

animals used for experimental and other scientific purposes). All strains

used in this study were bred into C57BL/6 background. The day of vaginal

plug was considered to be E0.5. Analysis of the phenotype and genotype

were performed independently in blind experiments and compared afterward.

Themouse ESC line XB814 harboring a gene trap insertion in Jmjd3was ob-

tained from BayGenomics (http://www.mmrrc.org/catalog/overview_BG.php)

(Stryke et al., 2003) and the insertion site of the trap cassette was mapped to

Jmjd3 intron 1 by sequence analysis. ESCswere injected into C57BL/6 blasto-

cysts to generate chimeras following standard procedures (Nagy et al., 2003).

Germline transmission was confirmed by PCR on genomic DNA extracted

from tail biopsies using the primers Forw1, Rev1, and Rev2 (Table S1). The

Rosa26flSTOP-Jmjd3 and BAC transgenic Jmjd3mut strains are described in

Extended Experimental Procedures.

Immunohistochemistry and 3D Reconstruction

Detailed immunohistochemistry procedures are described in the Extended

Experimental Procedures. For 3D reconstructions we used the software

Reconstruct (Fiala, 2005). Consecutive coronal sections (each 8 mm thick)

from brainstems of E18.5 Jmjd3+/+ and Jmjd3�/� embryos were stained for

NK1R and SST. Images were aligned manually and NK1R-positive cells

were defined using the Wildfire tool with the same threshold settings for all

images. 3D visualization was generated by stacking the traces defined in

each section.

Respiratory In Vivo and In Vitro Studies

As previously reported (Blanchi et al., 2003; Caubit et al., 2010; Ren et al.,

2003), in vivo and in vitro approaches were used to examine the respiratory

system of mouse embryos after surgical delivery at gestational days E15.5,

E16.5, and E18.5. At E18.5, whole-body plethysmography (EMKA technolo-

gies, Paris, France) was used to record in vivo breathing movements for

3–5 min, just after delivery. Plethysmographic chambers were maintained at

32�C and embryonic mouth temperature monitored with miniature thermistor

nylon-coated probe. Thin cooper wires inserted within intercostal muscles

allowed recordings of chest EMG and ECG in either E18.5 embryos kept in

utero with preserved umbilical irrigation or younger embryos studied in vitro.

For in vitro experiments, the medulla and cervical cord of E18.5 embryos

were dissected, placed in a 2 ml chamber, superfused with carbogenated

aCSF (4 ml/min, 26�C [pH 7.4]) and the C4 root of the phrenic nerve was

sucked within a glass micropipette to record rhythmic PBs. In younger

embryos where the C4 root was especially fragile, the rib cage was retained

in vitro to allow visualization of chest contractions and recording of chest respi-

ratory EMG. Control aCSF was occasionally replaced by acidified aCSF (pH

7.1), aCSF containing drugs or aCSFwith elevated concentration of potassium

for 5–10min. Electrical stimulations were applied via tungstenmicroelectrodes

to the diaphragm or phrenic nerve to check the embryonic neuromuscular

system and to the spinal cord or medulla to check the neural excitability and

synaptic transmission of the embryonic medullospinal pathways. In addition,

brainstems of E16.5 embryos were dissected and used for HPLC biochemical

analysis of serotonin metabolism (Menuet et al., 2011).

SUPPLEMENTAL INFORMATION

Supplemental Information includes Extended Experimental Procedures, five

figures, one table, and one movie and can be found with this article online at

http://dx.doi.org/10.1016/j.celrep.2012.09.013.

LICENSING INFORMATION

This is an open-access article distributed under the terms of the Creative

Commons Attribution-Noncommercial-No Derivative Works 3.0 Unported

License (CC-BY-NC-ND; http://creativecommons.org/licenses/by-nc-nd/3.

0/legalcode).

ACKNOWLEDGMENTS

We thank Alastair D. Morrison and Philip D. Hayes (GlaxoSmithKline Research

and Development), and Sigrun Murr and Elisa Allievi (IFOM-IEO Campus

Transgenic facility) for technical help and support. This work was supported

by grants from the Italian Association for Cancer Research (AIRC) (G.T.,

S.C., and G.N.), the Italian Health Ministry (G.T. and S.C.), the Italian Founda-

tion for Cancer Research (FIRC) (P.P.T., G.T., and S.C.), the Association for

International Cancer Research (AICR) (G.T.), the Italian Ministry for Education,

University and Research (S.C.) and the Centre National de la Recherche

Scientifique (UMR 6231, G.H.). G.T. conceived the study; T.B. generated the

Jmjd3 knockout strain and performed the experiments along with N.V.,

C.M., M.B., C.G., and G.H. who performed the electrophysiological character-

ization; M.C. and S.C. generated the Rosa26flSTOP-Jmjd3 strain; P.P.T. gener-

ated and characterized the Jmjd3mut BAC transgenic strains; B.K.T., F.S.,

S.B., G.N., and L.K. provided reagents and helped with the characterization

of cells and mice; A.S. performed in situ hybridizations; T.B., N.V., G.H., and

G.T. analyzed the data; G.T. wrote the paper.

Received: February 5, 2012

Revised: July 10, 2012

Accepted: September 12, 2012

Published: October 25, 2012

REFERENCES

Agger, K., Cloos, P.A., Christensen, J., Pasini, D., Rose, S., Rappsilber, J., Is-

saeva, I., Canaani, E., Salcini, A.E., and Helin, K. (2007). UTX and JMJD3 are

histone H3K27 demethylases involved in HOX gene regulation and develop-

ment. Nature 449, 731–734.

Akizu, N., Estaras, C., Guerrero, L., Martı, E., and Martınez-Balbas, M.A.

(2010). H3K27me3 regulates BMP activity in developing spinal cord. Develop-

ment 137, 2915–2925.

Amiel, J., Dubreuil, V., Ramanantsoa, N., Fortin, G., Gallego, J., Brunet, J.F.,

and Goridis, C. (2009). PHOX2B in respiratory control: lessons from congenital

central hypoventilation syndrome and its mousemodels. Respir. Physiol. Neu-

robiol. 168, 125–132.

Bibel, M., Richter, J., Schrenk, K., Tucker, K.L., Staiger, V., Korte, M., Goetz,

M., and Barde, Y.A. (2004). Differentiation of mouse embryonic stem cells

into a defined neuronal lineage. Nat. Neurosci. 7, 1003–1009.

Blanchi, B., Kelly, L.M., Viemari, J.C., Lafon, I., Burnet, H., Bevengut, M., Till-

manns, S., Daniel, L., Graf, T., Hilaire, G., and Sieweke,M.H. (2003). MafB defi-

ciency causes defective respiratory rhythmogenesis and fatal central apnea at

birth. Nat. Neurosci. 6, 1091–1100.

Bou-Flores, C., Lajard, A.M., Monteau, R., De Maeyer, E., Seif, I., Lanoir, J.,

and Hilaire, G. (2000). Abnormal phrenic motoneuron activity and morphology

in neonatal monoamine oxidase A-deficient transgenic mice: possible role of

a serotonin excess. J. Neurosci. 20, 4646–4656.

Bouvier, J., Thoby-Brisson, M., Renier, N., Dubreuil, V., Ericson, J., Champag-

nat, J., Pierani, A., Chedotal, A., and Fortin, G. (2010). Hindbrain interneurons

and axon guidance signaling critical for breathing. Nat. Neurosci. 13, 1066–

1074.

Burgold, T., Spreafico, F., De Santa, F., Totaro, M.G., Prosperini, E., Natoli, G.,

and Testa, G. (2008). The histone H3 lysine 27-specific demethylase Jmjd3 is

required for neural commitment. PLoS ONE 3, e3034.

Byvoet, P., Shepherd, G.R., Hardin, J.M., and Noland, B.J. (1972). The distri-

bution and turnover of labeled methyl groups in histone fractions of cultured

mammalian cells. Arch. Biochem. Biophys. 148, 558–567.

Cell Reports 2, 1–15, October 25, 2012 ª2012 The Authors 13

Please cite this article in press as: Burgold et al., The H3K27 Demethylase JMJD3 Is Required for Maintenance of the Embryonic Respiratory NeuronalNetwork, Neonatal Breathing, and Survival, Cell Reports (2012), http://dx.doi.org/10.1016/j.celrep.2012.09.013

Canovas, S., Cibelli, J.B., and Ross, P.J. (2012). Jumonji domain-containing

protein 3 regulates histone 3 lysine 27 methylation during bovine preimplanta-

tion development. Proc. Natl. Acad. Sci. USA 109, 2400–2405.

Caubit, X., Thoby-Brisson, M., Voituron, N., Filippi, P., Bevengut, M., Faralli,

H., Zanella, S., Fortin, G., Hilaire, G., and Fasano, L. (2010). Teashirt 3 regulates

development of neurons involved in both respiratory rhythm and airflow

control. J. Neurosci. 30, 9465–9476.

Chen, S., Ma, J., Wu, F., Xiong, L.J., Ma, H., Xu, W., Lv, R., Li, X., Villen, J.,

Gygi, S.P., et al. (2012). The histone H3 Lys 27 demethylase JMJD3 regulates

gene expression by impacting transcriptional elongation. Genes Dev. 26,

1364–1375.

De Santa, F., Totaro, M.G., Prosperini, E., Notarbartolo, S., Testa, G., and Na-

toli, G. (2007). The histone H3 lysine-27 demethylase Jmjd3 links inflammation

to inhibition of polycomb-mediated gene silencing. Cell 130, 1083–1094.

De Santa, F., Narang, V., Yap, Z.H., Tusi, B.K., Burgold, T., Austenaa, L., Bucci,

G., Caganova, M., Notarbartolo, S., Casola, S., et al. (2009). Jmjd3 contributes

to the control of gene expression in LPS-activated macrophages. EMBO J. 28,

3341–3352.

Fiala, J.C. (2005). Reconstruct: a free editor for serial section microscopy.

J. Microsc. 218, 52–61.

Fortin, G., and Thoby-Brisson, M. (2009). Embryonic emergence of the respi-

ratory rhythm generator. Respir. Physiol. Neurobiol. 168, 86–91.

Gaspar, P., Cases, O., and Maroteaux, L. (2003). The developmental role

of serotonin: news from mouse molecular genetics. Nat. Rev. Neurosci. 4,

1002–1012.

Gestreau, C., Heitzmann, D., Thomas, J., Dubreuil, V., Bandulik, S., Reichold,

M., Bendahhou, S., Pierson, P., Sterner, C., Peyronnet-Roux, J., et al. (2010).

Task2 potassium channels set central respiratory CO2 and O2 sensitivity.

Proc. Natl. Acad. Sci. USA 107, 2325–2330.

Glaser, S., Schaft, J., Lubitz, S., Vintersten, K., van der Hoeven, F., Tufteland,

K.R., Aasland, R., Anastassiadis, K., Ang, S.L., and Stewart, A.F. (2006).

Multiple epigenetic maintenance factors implicated by the loss of Mll2 in

mouse development. Development 133, 1423–1432.

Gray, P.A., Janczewski, W.A., Mellen, N., McCrimmon, D.R., and Feldman,

J.L. (2001). Normal breathing requires preBotzinger complex neurokinin-1

receptor-expressing neurons. Nat. Neurosci. 4, 927–930.

Gray, P.A., Fu, H., Luo, P., Zhao, Q., Yu, J., Ferrari, A., Tenzen, T., Yuk, D.I.,

Tsung, E.F., Cai, Z., et al. (2004). Mouse brain organization revealed through

direct genome-scale TF expression analysis. Science 306, 2255–2257.

Gray, P.A., Hayes, J.A., Ling, G.Y., Llona, I., Tupal, S., Picardo, M.C., Ross,

S.E., Hirata, T., Corbin, J.G., Eugenın, J., and Del Negro, C.A. (2010). Develop-

mental origin of preBotzinger complex respiratory neurons. J. Neurosci. 30,

14883–14895.

Hansen, K.H., Bracken, A.P., Pasini, D., Dietrich, N., Gehani, S.S., Monrad, A.,

Rappsilber, J., Lerdrup, M., and Helin, K. (2008). A model for transmission of

the H3K27me3 epigenetic mark. Nat. Cell Biol. 10, 1291–1300.

Hilaire, G., Voituron, N., Menuet, C., Ichiyama, R.M., Subramanian, H.H., and

Dutschmann, M. (2010). The role of serotonin in respiratory function and

dysfunction. Respir. Physiol. Neurobiol. 174, 76–88.

Hirabayashi, Y., Suzki, N., Tsuboi, M., Endo, T.A., Toyoda, T., Shinga, J., Ko-

seki, H., Vidal, M., and Gotoh, Y. (2009). Polycomb limits the neurogenic

competence of neural precursor cells to promote astrogenic fate transition.

Neuron 63, 600–613.

Hunt, P.N., McCabe, A.K., and Bosma, M.M. (2005). Midline serotonergic neu-

rones contribute to widespread synchronized activity in embryonic mouse

hindbrain. J. Physiol. 566, 807–819.

Jepsen, K., Solum, D., Zhou, T., McEvilly, R.J., Kim, H.J., Glass, C.K.,

Hermanson, O., and Rosenfeld, M.G. (2007). SMRT-mediated repression of

an H3K27 demethylase in progression from neural stem cell to neuron. Nature

450, 415–419.

Kennison, J.A. (1995). The Polycomb and trithorax group proteins of

Drosophila: trans-regulators of homeotic gene function. Annu. Rev. Genet.

29, 289–303.

14 Cell Reports 2, 1–15, October 25, 2012 ª2012 The Authors

Lan, F., Bayliss, P.E., Rinn, J.L., Whetstine, J.R., Wang, J.K., Chen, S., Iwase,

S., Alpatov, R., Issaeva, I., Canaani, E., et al. (2007). A histone H3 lysine 27 de-

methylase regulates animal posterior development. Nature 449, 689–694.

Margueron, R., Justin, N., Ohno, K., Sharpe, M.L., Son, J., Drury, W.J., 3rd,

Voigt, P., Martin, S.R., Taylor, W.R., De Marco, V., et al. (2009). Role of the

polycomb protein EED in the propagation of repressive histone marks. Nature

461, 762–767.

Menuet, C., Borghgraef, P., Matarazzo, V., Gielis, L., Lajard, A.M., Voituron, N.,

Gestreau, C., Dutschmann, M., Van Leuven, F., and Hilaire, G. (2011). Raphe

tauopathy alters serotonin metabolism and breathing activity in terminal

Tau.P301L mice: possible implications for tauopathies and Alzheimer’s

disease. Respir. Physiol. Neurobiol. 178, 290–303.

Miller, S.A., Mohn, S.E., andWeinmann, A.S. (2010). Jmjd3 and UTX play a de-

methylase-independent role in chromatin remodeling to regulate T-box family

member-dependent gene expression. Mol. Cell 40, 594–605.

Mohn, F., Weber, M., Rebhan, M., Roloff, T.C., Richter, J., Stadler, M.B., Bibel,

M., and Schubeler, D. (2008). Lineage-specific polycomb targets and de novo

DNA methylation define restriction and potential of neuronal progenitors. Mol.

Cell 30, 755–766.

Nagy, A., Gertsenstein, M., Vintersten, K., and Behringer, R. (2003). Manipu-

lating the Mouse Embryo: A Laboratory Manual, Third Edition (Cold Spring

Harbor, NY: Cold Spring Harbor Laboratory Press).

Natoli, G., Testa, G., and De Santa, F. (2009). The future therapeutic potential

of histone demethylases: A critical analysis. Curr. Opin. Drug Discov. Devel.

12, 607–615.

Nishida, K., Nakayama, K., Yoshimura, S., and Murakami, F. (2011). Role of

Neph2 in pontine nuclei formation in the developing hindbrain. Mol. Cell. Neu-

rosci. 46, 662–670.

Onimaru, H., Ikeda, K., and Kawakami, K. (2009). Phox2b, RTN/pFRG neurons

and respiratory rhythmogenesis. Respir. Physiol. Neurobiol. 168, 13–18.

Pereira, J.D., Sansom, S.N., Smith, J., Dobenecker, M.W., Tarakhovsky, A.,

and Livesey, F.J. (2010). Ezh2, the histone methyltransferase of PRC2, regu-

lates the balance between self-renewal and differentiation in the cerebral

cortex. Proc. Natl. Acad. Sci. USA 107, 15957–15962.

Ren, J., Lee, S., Pagliardini, S., Gerard, M., Stewart, C.L., Greer, J.J., andWev-

rick, R. (2003). Absence of Ndn, encoding the Prader-Willi syndrome-deleted

gene necdin, results in congenital deficiency of central respiratory drive in

neonatal mice. J. Neurosci. 23, 1569–1573.

Rose,M.F., Ren, J., Ahmad, K.A., Chao, H.T., Klisch, T.J., Flora, A., Greer, J.J.,

and Zoghbi, H.Y. (2009). Math1 is essential for the development of hindbrain

neurons critical for perinatal breathing. Neuron 64, 341–354.

Satoh, T., Takeuchi, O., Vandenbon, A., Yasuda, K., Tanaka, Y., Kumagai, Y.,

Miyake, T., Matsushita, K., Okazaki, T., Saitoh, T., et al. (2010). The Jmjd3-Irf4

axis regulates M2 macrophage polarization and host responses against

helminth infection. Nat. Immunol. 11, 936–944.

Smith, J.C., Ellenberger, H.H., Ballanyi, K., Richter, D.W., and Feldman, J.L.

(1991). Pre-Botzinger complex: a brainstem region that may generate respira-

tory rhythm in mammals. Science 254, 726–729.

Solomon, I.C., Edelman, N.H., and Neubauer, J.A. (2000). Pre-Botzinger

complex functions as a central hypoxia chemosensor for respiration in vivo.

J. Neurophysiol. 83, 2854–2868.

Stornetta, R.L., Rosin, D.L., Wang, H., Sevigny, C.P., Weston, M.C., and Guye-

net, P.G. (2003). A group of glutamatergic interneurons expressing high levels

of both neurokinin-1 receptors and somatostatin identifies the region of the

pre-Botzinger complex. J. Comp. Neurol. 455, 499–512.

Stryke, D., Kawamoto, M., Huang, C.C., Johns, S.J., King, L.A., Harper, C.A.,

Meng, E.C., Lee, R.E., Yee, A., L’Italien, L., et al. (2003). BayGenomics:

a resource of insertional mutations in mouse embryonic stem cells. Nucleic

Acids Res. 31, 278–281.

Tan, W., Sherman, D., Turesson, J., Shao, X.M., Janczewski, W.A., and Feld-

man, J.L. (2012). Reelin demarcates a subset of pre-Botzinger complex

neurons in adult rat. J. Comp. Neurol. 520, 606–619.

Please cite this article in press as: Burgold et al., The H3K27 Demethylase JMJD3 Is Required for Maintenance of the Embryonic Respiratory NeuronalNetwork, Neonatal Breathing, and Survival, Cell Reports (2012), http://dx.doi.org/10.1016/j.celrep.2012.09.013

Testa, G. (2011). The time of timing: how Polycomb proteins regulate neuro-

genesis. Bioessays 33, 519–528.

Thoby-Brisson, M., Karlen, M., Wu, N., Charnay, P., Champagnat, J., and For-

tin, G. (2009). Genetic identification of an embryonic parafacial oscillator

coupling to the preBotzinger complex. Nat. Neurosci. 12, 1028–1035.

Thoby-Brisson, M., Trinh, J.B., Champagnat, J., and Fortin, G. (2005). Emer-

gence of the pre-Botzinger respiratory rhythm generator in the mouse embryo.

J. Neurosci. 25, 4307–4318.

Viemari, J.C., Bevengut, M., Burnet, H., Coulon, P., Pequignot, J.M., Tiveron,

M.C., and Hilaire, G. (2004). Phox2a gene, A6 neurons, and noradrenaline are

essential for development of normal respiratory rhythm in mice. J. Neurosci.

24, 928–937.

Wallen-Mackenzie, A., Gezelius, H., Thoby-Brisson, M., Nygard, A., Enjin, A.,

Fujiyama, F., Fortin, G., and Kullander, K. (2006). Vesicular glutamate trans-

porter 2 is required for central respiratory rhythm generation but not for loco-

motor central pattern generation. J. Neurosci. 26, 12294–12307.

Xiang, Y., Zhu, Z., Han, G., Lin, H., Xu, L., and Chen, C.D. (2007). JMJD3 is

a histone H3K27 demethylase. Cell Res. 17, 850–857.

Yu, B.D., Hanson, R.D., Hess, J.L., Horning, S.E., and Korsmeyer, S.J. (1998).

MLL, amammalian trithorax-group gene, functions as a transcriptional mainte-

nance factor in morphogenesis. Proc. Natl. Acad. Sci. USA 95, 10632–10636.

Zanella, S., Watrin, F., Mebarek, S., Marly, F., Roussel, M., Gire, C., Diene, G.,

Tauber, M., Muscatelli, F., and Hilaire, G. (2008). Necdin plays a role in the

serotonergic modulation of the mouse respiratory network: implication for

Prader-Willi syndrome. J. Neurosci. 28, 1745–1755.

Cell Reports 2, 1–15, October 25, 2012 ª2012 The Authors 15

Supplemental Information

EXTENDED EXPERIMENTAL PROCEDURES

Generation of Rosa26flSTOP-Jmjd3 MiceTo generate tissue-specific conditional Jmjd3 transgenic mice, we cloned full-length mouse Jmjd3 cDNA into the pR26 loxP2STOP

frt2 IRES eGFP targeting vector (Sasaki et al., 2006). C57BL/6-derived ESC were transfected, cultured and selected as previously

described (Casola, 2004). Correctly targeted ESCswere injected into C57BL/6 albino blastocysts to generate chimeras and germline

transmission was confirmed by Southern blot (Sasaki et al., 2006).

Expression of the Jmjd3 transgene was induced upon breeding to the ubiquitous Cre-deleter strain PGK-Cre (Lallemand et al.,

1998).

For PCR genotyping of genomic DNA extracted from tail biopsies the primers Forw2, Rev3 and Rev4 were used (Table S1).

Generation of BAC Transgenic Jmjd3mut MiceAmouse BAC (RP23-107C10) containing the Jmjd3 gene locus was engineered to insert a point mutation resulting in the Histidine to

Alanine substitution at position 1388 (H1388A). Microinjection into fertilized oocytes and generation of transgenic mice was per-

formed by standard techniques using oocytes derived from FVB mice (Muyrers et al., 2004; Testa et al., 2004; Vintersten et al.,

2004). Transgenic founder animals were selected by PCR of tail biopsies using primer pair FGt and RGt (Table S1). Two transgenic

founders (Jmjd3mut21 and Jmjd3mut65) were selected for further characterization and the number of BAC integrations was assessed

by Southern blot and real-time PCR using a TaqMan copy number assay. In order to confirm the expression of themutant Jmjd3 allele

brain cDNA of Jmjd3mut21 and Jmjd3mut65 transgenic mice was amplified by PCR using the primers FRTseq and RRTseq (Table S1).

PCR products were subcloned into pCR2.1-TOPO vector (Invitrogen) and sequenced.

In order to distinguish the Jmjd3 trap allele from the Jmjd3mutBAC transgene in compound heterozygous crosses, a TaqMan copy

number assay was designed based on a probe spanning the junction between intron 1 and the gene trap cassette (ProbecnTrapCACTTTCGCCCCCTCAGCCAG) along with the amplification with primers FcnTrap and RcnTrap (Table S1).

ESC Culture, Derivation, and DifferentiationESC line XB814 (BayGenomics) was propagated in feeder-free conditions, at 37�C, 5% CO2 in a humidified atmosphere with daily

medium change. Standard ESCmedium consisted of DMEMmedium (with high glucose, with sodium pyruvate, without L-glutamine)

supplemented with 1000 U/ml leukemia inhibitory factor (LIF; Chemicon), 15% FBS (tested for ESC culture; GIBCO), 2 mM L-gluta-

mine, 0.1 mM NEAA, 25 mM HEPES and 0.1 mM 2-mercaptoethanol.

To derive ESCs homozygous for the Jmjd3 trap allele, timed matings between heterozygous XB814 mice were set up. Plugged

females were sacrificed at 3.5 days postcoitus (dpc) and uteri were flushed to collect the embryos at the morula stage. They were

cultured in vitro in drop cultures of KSOM medium (95 mM NaCl, 2.5 mM KCl, 0.35 mM KH2PO4, 0.20 mM MgSO4 7H20,

0.20 mM glucose, 10 mM sodium lactate, 25 mM NaHCO3 phenol red, 0.2 mM sodium pyruvate, 1.71 mM CaCl2 2H20, 0.01 mM

EDTA, 1 mM L-glutamine and 1 mg/ml BSA) covered in embryo-tested mineral oil (Sigma) for 24 hr in a humidified atmosphere of

5% CO2 at 37�C. Blastocysts were transferred individually to 48-well plate wells containing mitotically inactivated mouse embryonic

fibroblasts (MEFs). The medium for ESC derivation consisted of standard ESCmedium supplemented with 50 mMmitogen-activated

protein kinase/extracellular signal-regulated kinase (MEK1) inhibitor PD98059 (Cell Signaling) (Buehr and Smith, 2003; Burdon et al.,

1999). After five to six days, blastocyst outgrowths were disaggregated into small clumps in trypsin-EDTA complemented with 1%

chicken serum. Cells were replated onto a 24-well plate with inactivated MEFs and allowed to grow for three to five days. After this

period, the cells were no longer cultured in the presence of the MEK inhibitor and were progressively expanded onto larger surfaces.

ESC differentiation into glutamatergic neurons was carried out as described (Bibel et al., 2007)

RNA Extraction and cDNA SynthesisTotal RNA was isolated with TRIZOL Reagent (Invitrogen) for ESC lines or RNeasy Mini Kit (QIAGEN) for fetal brains according to the

manufacturer’s instructions. cDNA synthesis was carried out with the SuperScript VILO cDNA Synthesis Kit (Invitrogen) following

manufacturer’s instructions.

Gene Expression AnalysisqRT-PCR was performed on 7900HT Fast Real-Time PCR system (Applied Biosystems) using TaqMan Gene expression assays

(Applied Biosystems): Jmjd3 (Mm01332680_m1), Tbp (Mm00446973_m1), Esrrg (Mm00516267_m1), Kirrel3 (Mm00659293_m1),

March4 (Mm03015817_m1) andReelin (Mm00465200_m1). Each sample was analyzed in triplicates and normalized to TATA-Binding

Protein (Tbp) used as normalization control. RelativemRNA amounts were calculated by the comparative cycle threshold (Ct) method

using the 2-DCt formula.

Copy Number AssayGene copy numbers were determined by real-time PCR using TaqMan copy number assays (Applied Biosystems): Jmjd3

(Mm00351536_cn). Each sample was analyzed in quadruplicates and normalized to the reference gene Telomerase reverse tran-

scriptase (TERT). Absolute copy numbers were determined by relative quantitation using the comparative (2-DDCt) method.

Cell Reports 2, 1–15, October 25, 2012 ª2012 The Authors S1

RT-PCRIn order to identify the Jmjd3 transcript variants we used the following primers: Fex.1-2/Rex.1-2, Fex.2-3/Rex.2-3, Fex.3-4/Rex.3-4, Fex.4-5/

Rex.4-5, Fex.14-16/Rex.14-16, FTbp/RTbp (Table S1).

Southern BlottingGenomic DNA was digested with the relevant enzyme at 37�C overnight and the entire digest was run on an 1 3 TAE agarose gel.

DNA was transferred onto a Hybond-N+ nylon membrane (GE Healthcare) in alkaline transfer buffer overnight. Filter was first prehy-

bridized in modified Church and Gilbert buffer (0.5 M sodium phosphate buffer pH 7.2, 7% SDS, 1 mM EDTA) for at least 30 min at

65�C and then hybridized in the same buffer upon addition of the relevant DNA probe at 65�C overnight. Radioactively labeled DNA

probes were prepared by random primed labeling with the Ladderman Labeling kit (Takara), according to themanufacturer’s instruc-

tions. Autoradiography was performed by exposing the filter to a storage phosphor screen and subsequent signal detection using

a Typhoon scanner system (GE Healthcare).

In Situ HybridizationBrains were dissected in a physiological solution of 0.9% NaCl and 0.6% glucose, fixed overnight in 4% paraformaldehyde (PFA) at

4�C, cryoprotected in 30% sucrose overnight at 4�C and embedded in cryostat freezing medium (Bio-Optica). Brainstems were

sectioned at 8 mm and placed serially onto SuperFrost Ultra Plus glass slides (Thermo Fisher Scientific). Tissue sections were

post-fixed for 10 min in 4% PFA, permeabilized with 1 mg/ml proteinase K for 5 min, refixed in 4% PFA for 5 min, acetylated

for 10 min in 0.1 M triethanolamine-HCl with 0.25% acetic anhydride and blocked in hybridization buffer (50% deionized formamide,

53 SSC, 53 Denhardt’s Solution, 100 mg/ml yeast tRNA, 100 mg/ml denatured salmon sperm DNA) for 2 hr at RT. For hybridization

slides were incubated in hybridization buffer containing 500-700 ng/ml of digoxigenin (DIG)-labeled RNA probe (Roche) at 68�C over-

night. Slides were washed in SSC buffers and incubated with a sheep polyclonal anti-digoxigenin antibody conjugated to alkaline

phosphatase (Roche #11093274910, 1:2000) at 4�C overnight. The colorimetric staining was developed by incubation in BM Purple

(Roche), supplemented with 2 mM Levamisole. Slides were mounted with VectaMount AQ aqueous mounting medium (Vector Labo-

ratories) and staining was examined on an Olympus microscope equipped with a Nikon color camera.

In situ hybridization experiments using radiolabelled riboprobes were performed as described (Simeone, 1999).

The Jmjd3 antisense probe was synthesized by SP6 in vitro transcription of a 590 bp fragment in the 30 region of the Jmjd3 cDNA

comprising exons 19 to 23.

HistologyHistological analysis of lung sections was performed through standard hematoxylin/eosin staining by the Toronto Centre for Pheno-

genomics at the Centre for Modeling Human Disease (CMHD) Pathology Core.

ImmunohistochemistryBrains were dissected in a physiological solution of 0.9%NaCl and 0.6% glucose, fixed overnight in 4%PFA at 4�C, cryoprotected in

30% sucrose overnight at 4�C and embedded in cryostat freezing medium (Bio-Optica). Brainstems were sectioned at 8 mm and

placed serially onto SuperFrost Ultra Plus glass slides (Thermo Fisher Scientific). Sections were post-fixed for 10 min with 4%

PFA. For the detection of nuclear antigens an antigen enhancement step was performed by briefly boiling the slides in 10mM sodium

citrate pH 6.0 using a microwave. Sections were incubated in permeabilization/blocking solution (0.2% Triton X-100, 5% donkey

serum in PBS) for 30 min at RT and incubated with primary antibodies diluted in PBS plus 2% donkey serum, 0.05% Triton X-100

at 4�C overnight. The antibodies and their dilutions used in this study are: anti-NK1R (Sigma-Aldrich S8305, 1:5,000), anti-SST (Santa

Cruz Biotechnology sc-7819, 1:500), anti-PHOX2b (gift from C. Goridis, 1:500), anti-H3K27me3 (Cell Signaling #9733, 1:200) and

anti-VGLUT2 (Synaptic Systems 135403, 1:250). Specimens were incubated with species specific secondary antibodies conjugated

to Alexa488 (Invitrogen) and Cy3 (Jackson ImmunoResearch) and Dapi (Sigma-Aldrich) for 3 hr at RT and coverslipped with Vecta-

Mount AQ aqueous mounting medium (Vector Laboratories). Images were acquired with an Olympus fluorescence microscope

equipped with a CoolSNAP EZ CCD camera (Photometrics) using MetaVue software (Molecular Devices). For clarity, images were

adjusted for contrast and brightness using ImageJ.

Western BlotCells were lysed with 50 mM Tris-HCl pH 7.6, 0.15 M NaCl, 5 mM EDTA, 1% Triton X-100 supplemented with 13 protease inhibitor

cocktail (Roche) and samples were sonicated with a Diagenode Bioruptor sonicator. Proteins were separated on a 5% polyacryl-

amide gel and then transferred onto a PVDF Immobilon-P membrane (Millipore) in semi-dry conditions using a Trans-Blot SD.

Semi-Dry Transfer Cell (BioRad) at 90 mA for 2 hr. Membranes were blocked in 5% skimmed milk powder in Tris-buffered saline

with 0.1% Tween 20 (TBST) for 1 hr at RT. Membranes were incubated with the following antibodies, diluted in blocking solution,

at 4�C overnight: polyclonal anti-JMJD3 antibody (in house polyclonal antibody, described in (De Santa et al., 2007)) andmonoclonal

anti-VINCULIN (Sigma-Aldrich). Blots were incubated with secondary antibodies goat-anti-mouse IgG or goat anti-rabbit IgG conju-

gatedwith horseradish peroxidase (Bio-Rad) diluted in blocking solution for 30min at RT. Bandswere detected using AmershamECL

Western Blotting Detection Reagents (GE Healthcare) following the manufacturer’s instructions.

S2 Cell Reports 2, 1–15, October 25, 2012 ª2012 The Authors

Chromatin ImmunoprecipitationCells were cross-linked in 1% formaldehyde for 10 min at room temperature. Fixed samples were washed three times with PBS and

resuspended in SDS buffer (0.5% SDS, 50 mM Tris-Cl pH 8.1, 100 mM NaCl, 5 mM EDTA pH 8, 0.02% NaN3 diluted in water). Cells

were centrifuged and resuspended in ice-cold immunoprecipitation (IP) buffer (3:1 SDS buffer/Triton dilution buffer; Triton dilution

buffer: 5% Triton X-100, 100 mM Tris-Cl pH 8.6, 100 mM NaCl, 5 mM EDTA pH 8, 0.02% NaN3 diluted in water). Lysed samples

were sonicated and chromatin was quantified. For immunoprecipitation with H3 and H3K27me3 antibodies (rabbit monoclonal

from Cell Signaling) 70 mg of chromatin were used. For imunoprecipitation with Jmjd3 (in house polyclonal antibody, described in

De Santa et al., 2007) 500 mg of chromatin were used. For each immunoprecipitation chromatin was diluted in 1 ml IP buffer and

10 ml were taken as 1% of input. Primary antibodies were incubated overnight at 4�C. Each sample was incubated with protein

A-Sepharose beads (Amersham) for 4 hr, washed three times in 150 mM wash buffer and once in 500 mM wash buffer (1%

Triton-X, 150 mM or 500 mM NaCl, 20 mM Tris-Cl pH 8.0, 0.1% SDS, 2 mM EDTA pH 8 diluted in water). Chromatin was de-cross-

linked in 0.1%SDS, 0.1MNaHCO3 buffer at 65�Covernight. DNAwas purified and eluted in 100 ml of water. 4 ml of DNAwere used for

qRT-PCR reaction using the primers listed in Table S1. The abundance of target genomeDNAwas normalized relative to that of input.

Statistical AnalysisStatistical significance was determined by two-tailed Student’s t test. Breeding outcome was evaluated by Chi-square test.

SUPPLEMENTAL REFERENCES

Bibel, M., Richter, J., Lacroix, E., and Barde, Y.A. (2007). Generation of a defined and uniform population of CNS progenitors and neurons frommouse embryonic

stem cells. Nat. Protoc. 2, 1034–1043.

Buehr, M., and Smith, A. (2003). Genesis of embryonic stem cells. Philos. Trans. R. Soc. Lond. B Biol. Sci. 358, 1397–1402, discussion 1402.

Burdon, T., Chambers, I., Stracey, C., Niwa, H., and Smith, A. (1999). Signaling mechanisms regulating self-renewal and differentiation of pluripotent embryonic

stem cells. Cells Tissues Organs (Print) 165, 131–143.

Casola, S. (2004). Conditional gene mutagenesis in B-lineage cells. Methods Mol. Biol. 271, 91–109.

De Santa, F., Totaro, M.G., Prosperini, E., Notarbartolo, S., Testa, G., and Natoli, G. (2007). The histone H3 lysine-27 demethylase Jmjd3 links inflammation to

inhibition of polycomb-mediated gene silencing. Cell 130, 1083–1094.

Lallemand, Y., Luria, V., Haffner-Krausz, R., and Lonai, P. (1998). Maternally expressed PGK-Cre transgene as a tool for early and uniform activation of the Cre

site-specific recombinase. Transgenic Res. 7, 105–112.

Muyrers, J.P., Zhang, Y., Benes, V., Testa, G., Rientjes, J.M., and Stewart, A.F. (2004). ET recombination: DNA engineering using homologous recombination in

E. coli. Methods Mol. Biol. 256, 107–121.

Sasaki, Y., Derudder, E., Hobeika, E., Pelanda, R., Reth,M., Rajewsky, K., and Schmidt-Supprian,M. (2006). Canonical NF-kappaB activity, dispensable for B cell

development, replaces BAFF-receptor signals and promotes B cell proliferation upon activation. Immunity 24, 729–739.

Simeone, A. (1999). Detection of mRNA in tissue sections with radiolabelled riboprobes. In In situ hybridization. A practical approach, D.G.Wilkinson, ed. (Oxford:

Oxford University Press), pp. 69–86.

Testa, G., Vintersten, K., Zhang, Y., Benes, V., Muyrers, J.P., and Stewart, A.F. (2004). BAC engineering for the generation of ES cell-targeting constructs and

mouse transgenes. Methods Mol. Biol. 256, 123–139.

Vintersten, K., Testa, G., and Stewart, A.F. (2004). Microinjection of BAC DNA into the pronuclei of fertilized mouse oocytes. Methods Mol. Biol. 256, 141–158.

Cell Reports 2, 1–15, October 25, 2012 ª2012 The Authors S3

Figure S1. Normal Lung Histology in Jmjd3 Mutants, Related to Figure 2

H&E staining of lung sections fromwild-type (upper row) and Jmjd3mutant (lower row) neonates at P0. For each genotype two individual mouse lungs are shown.

S4 Cell Reports 2, 1–15, October 25, 2012 ª2012 The Authors

Figure S2. Normal Cardiac Activity in Jmjd3 Mutants, Related to Figure 3

Heart rate of E18.5 embryos was monitored in utero by electrocardiography. Bars represent the mean ± SEM of the numbers (n) of mice indicated on the x axis

below the graph.

Cell Reports 2, 1–15, October 25, 2012 ª2012 The Authors S5

Figure S3. Characterization of Respiratory Oscillators in Jmjd3�/� Embryos, Related to Figure 4

(A) Immunohistochemical staining for VGLUT2 on sagittal (first column) and coronal sections (second column) from wild-type (upper row) and Jmjd3�/� (lower

row) E18.5 brainstems.

(B) Immunohistochemical staining for SST (first column) and H3K27me3 (second column) on coronal sections from wild-type (upper row) and Jmjd3�/� (lower

row) E18.5 brainstems. A merged image of H3K27me3 (magenta), SST (green) and DAPI (gray) stainings is shown in the last column on the right.

(C) Immunohistochemical staining for PHOX2b (green, first column) and NK1R (magenta, second column) on sagittal sections fromwild-type and Jmjd3�/� E18.5

brainstems. Jmjd3+/+ and Jmjd3�/� embryos displayed immunoreactivity for PHOX2b andNK1R in the RTN/pFRG and for PHOX2b in the VII motor nucleus (white

dotted outline).

(D) Jmjd3 in situ hybridization on sagittal brain sections from wild-type and mutant E18.5 embryos. Jmjd3+/+ embryos displayed Jmjd3 expression in the ventral

respiratory column, including the VII motor nucleus and the PBC whereas in Jmjd3�/� mutants the signal was virtually undetectable.

(E) The two upper traces show rhythmic diaphragmatic EMG bursts (large signals) and ECG (small signals) recorded in vitro in Jmjd3+/+ and Jmjd3�/� E15.5

preparations retaining the rib cage. At E18.5, none of the Jmjd3�/� preparations produced rhythmic diaphragmatic bursts (lower trace, ECG).

Scale bars represent 100 mm. PBC: pre-Botzinger complex, RTN/pFRG: retrotrapezoid nucleus and parafacial respiratory group, SO: Superior olive, VII: facial

motor nucleus.

S6 Cell Reports 2, 1–15, October 25, 2012 ª2012 The Authors

Figure S4. Generation of a Conditional Jmjd3 Transgenic Allele, Related to Figure 5

Southern blot analysis of targeted ESC clones. Genomic DNA from wild-type (WT) and targeted ESC clones was digested with EcoRI, yielding a 15.4 kb band for

the Rosa26 wild-type allele and a 5.1 kb fragment for the targeted allele.

Cell Reports 2, 1–15, October 25, 2012 ª2012 The Authors S7

Figure S5. Recovery of Jmjd3�/�;Jmjd3mut21 and Jmjd3�/�;Jmjd3mut65 Compound Mutants at Embryonic Day E18.5, Related to Figure 6

Table showing the recovery of Jmjd3�/� mice carrying either the Jmjd3mut21 or the Jmjd3mut65 BAC transgene at E18.5. Numbers of embryos for the various

genotypes were measured separately for the Jmjd3mut21 and Jmjd3mut65 strains.

S8 Cell Reports 2, 1–15, October 25, 2012 ª2012 The Authors