Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2017 Sep 15.
Published in final edited form as: Dev Biol. 2016 Jun 28;417(2):139–157. doi: 10.1016/j.ydbio.2016.06.042

Mouse models of Hirschsprung Disease and other developmental disorders of the Enteric Nervous System: Old and new players

Nadege Bondurand a,b,, E Michelle Southard-Smith c,
PMCID: PMC5026931  NIHMSID: NIHMS806536  PMID: 27370713

Abstract

Hirschsprung disease (HSCR, intestinal aganglionosis) is a multigenic disorder with variable penetrance and severity that has a general population incidence of 1/5000 live births. Studies using animal models have contributed to our understanding of the developmental origins of HSCR and the genetic complexity of this disease. This review summarizes recent progress in understanding control of enteric nervous system (ENS) development through analyses in mouse models. An overview of signaling pathways that have long been known to control the migration, proliferation and differentiation of enteric neural progenitors into and along the developing gut is provided as a framework for the latest information on factors that influence enteric ganglia formation and maintenance. Newly identified genes and additional factors beyond discrete genes that contribute to ENS pathology including regulatory sequences, miRNAs and environmental factors are also introduced. Finally, because HSCR has become a paradigm for complex oligogenic diseases with non-Mendelian inheritance, the importance of gene interactions, modifier genes, and initial studies on genetic background effects are outlined.

Keywords: Hirschsprung, enteric nervous system, neural crest, mouse models

Introduction

Identifying the factors that lead to enteric ganglia deficiencies, such as Hirschsprung disease (HSCR), has been a long-term goal of many investigators who study innervation of the intestine. Efforts have been particularly focused on genetic factors and the molecular effects of discrete coding and non-coding mutations as well as post-translational mechanisms that impact the enteric neural crest cells (ENCCs) that populate the fetal intestine to form the mature enteric nervous system (ENS). Current knowledge has greatly benefited from exchanges between human geneticists and researchers using animal models. Because gene targeting techniques and inbred strains have long been available in laboratory mice, advances in identifying genes that contribute to HSCR susceptibility and subsequent studies of cellular mechanism have been possible through analysis of ENS development in mice. In recent years, the impact of environmental factors on ENS development has been identified and has been advanced by studies in mouse models (Fu et al., 2010; Heuckeroth and Schafer, 2016; Schill et al., 2016).

The ENS is an extensive network of neurons and glial cells within the wall of the bowel that controls gut motility, regulates transport of ions across the epithelium and modulates blood flow (Furness, 2012). In the small and large intestine, neurons and glial cells are mostly found in two main plexuses: the myenteric plexus located between the circular and longitudinal muscle layers, and the submucosal found within the connective tissue of the submucosa. The essential role of enteric neurons in peristalsis control is exemplified by bowel obstruction that occurs in aganglionic regions of patients presenting with HSCR (Chakravarti et al., 2004). This multigenic disorder exhibits variable penetrance and severity with a general population incidence of 1/5000 live births and a prominent gender bias of 4:1 in males compared to females (Badner et al., 1990; Spouge and Baird, 1985). Studies using animal models have contributed to our understanding of HSCR genetic complexity through genome targeting efforts in mice that have identified many causative genes for aganglionosis. Continued studies in mice with complementary work in chick and zebrafish, have identified many other molecules that are crucial for ENS development and have aided in understanding cellular processes that occur in normal and abnormal ENS development (see examples included in recent reviews (Goldstein et al., 2013; Harrison and Shepherd, 2013; Lake and Heuckeroth, 2013; Obermayr et al., 2013; Zimmer and Puri, 2015). As a result, we now know that that neural crest cells, mainly from vagal levels of the neural tube, enter the foregut and migrate to colonize the whole length of the intestine (Burns and Douarin, 1998; Le Douarin and Teillet, 1974; Young et al., 1998). The principal pathway from vagal levels occurs in a rostro-caudal wave down the length of the developing intestine. Upon reaching the hindgut ENCCs can proceed either through a trans-mesenteric pathway (Nishiyama et al., 2012) or migrate through the cecum to populate the distal colon (Druckenbrod and Epstein, 2005). Because the gut lengthens substantially while it is being colonized, vagal ENCCs migrate further than any other neural crest cell population. Vagal progenitors are complemented by truncal and sacral ENCC populations that make smaller contributions to total cell numbers (Burns and Douarin, 1998; Kapur, 2000; Wang et al., 2011). In humans, it takes three weeks for these cell populations to colonize the whole length of the bowel, while in mice it takes five days, which is ¼ of the gestation period (for recent review see (McKeown et al., 2013; Obermayr et al., 2013)). During the whole process, coordinated proliferation, migration, and differentiation is required, as perturbations to ENCCs number, migratory behavior or rate of differentiation can result in aganglionosis of the distal bowel. Although apoptosis is not prominent while ENCCs are colonizing the gut (Gianino et al., 2003), early cell death before vagal crest cells enter the gut has been reported in the chick (Wallace et al., 2009) and is also known to occur among mutants that later exhibit intestinal aganglionosis (Durbec et al., 1996; Kapur, 1999; Stanchina et al., 2006).

This review summarizes recent progress in understanding control of ENS development through analysis of mouse models. An overview of signaling pathways that have long been known to control the migration, proliferation and differentiation of ENCCs into and along the developing gut is provided as a framework for the latest information on factors that influence enteric ganglia formation and maintenance. Newly discovered genes that cause HSCR or other abnormalities of enteric ganglia density such as hypoganglionosis or hyperganglionosis are described. Beyond discrete gene identification, the role of regulatory sequences, miRNAs and environmental factors in the etiology of ENS disorders are also introduced. Finally because HSCR has become a paradigm for complex oligogenic diseases with non-Mendelian inheritance, the importance of gene interactions, modifier genes, and initial studies on genetic background effects are included.

Molecular mediators that control ENS development and maturation: Key historical genes and new players

Over the years, multiple naturally occurring (“spontaneous”) or gene-targeted mutations that alter molecules involved in the colonization of the gut by ENCCs have been described. These include factors secreted by the gut mesenchyme that act on receptors expressed by ENCCs, transcription factors, guidance factors and morphogens, as well as proteins that transmit signals from the cell surface to the cytoskeleton and the nucleus, including adhesion molecules. Mutations in genes encoding many of these components have been associated with HSCR in human patients, and the majority of these factors are known to affect multiple cellular processes during development.

Between 2012 and 2013, several reviews described in detail many of the identified molecules that are known to play key roles during ENS development (Bergeron et al., 2013; Bondurand and Sham, 2013; Butler Tjaden and Trainor, 2013; Goldstein et al., 2013; Harrison and Shepherd, 2013; Lake and Heuckeroth, 2013; McKeown et al., 2013; Musser and Southard-Smith, 2013; Obermayr et al., 2013; Young, 2012). For details concerning what we refer to here as “key historical genes “, we encourage readers to go back to these elegant reviews. In order to highlight new data published over the last five years, we provide here an overview of individual genes that cause aganglionosis and their roles in ENS development when known (Table 1). We complement this with a summary of known genes that impact the ENS although aganglionosis is not evident (Table 2) and a summary of environmental factors that influence ENS development (Table 3). Finally we include a summary of known genetic interactions that affect ENS development (Table 4).

Table 1.

Single Gene Mutations in Mice that Lead to Intestinal Aganglionosis and Their Effects on ENS Development.

Gene Allele Type Developmental
Effects on
ENS
Effect on Mature ENS Mouse
Model
References
Initial Gene
Report in
HSCR
patients*
Col6a4

Collagen 6α4;
Tg (Sox3-
GFP,Tyr)HolNpln

(aka “Holstein”)



Untargeted transgene
insertion upstream of
Col6a4 leads to up-
regulated Col6a4
expression
Delayed
colonization of
fetal intestine by
ENCCs.

Reduced
glial/progenitor
cell ratios.
Homozygotes exhibit
aganglionosis in distal
colon accompanied by
hypoganglionosis in mid
colon.
(Soret et al., 2015) Mutation not
reported in GI
patients
Ece1

Endothelin
converting
enzyme 1,
Processes
endothelins to
active
peptides
Ece1tm1Reh / tm1/Reh

(aka Ece1−/−)


Targeted gene knockout
Defective
migration of
ENCCs into the
developing
hindgut.
Aganglionosis in distal
colon
(Yanagisawa et al., 1998) (Hofstra et al., 1999)
Edn3


Endothelin 3
peptide,
Ednrb ligand,
Edn3ls / ls

(aka “Lethal
Spotting”)


spontaneous point mutation
Defective
migration of
ENCCs into the
developing
hindgut.
Aganglionosis in distal
colon; decreased neuronal
numbers
(Baynash et al., 1994;
Coventry et al., 1994;
Rothman and Gershon, 1984)
(Hofstra et al., 1996)
Edn3tm1Ywa/ tm1Ywa

(aka Edn3 −/−)


Targeted gene knockout
uncharacterized Aganglionosis in distal
colon
(Baynash et al., 1994)
Ednrb



Endothelin
receptor type
B, G-
coupled
Protein
Receptor
Ednrbs/s

(aka “piebald”)

spontaneous insertion in
intron causes reduced gene
expression.
uncharacterized Hypomorphic allele due to
decreased gene
expression. Very rare
colonic aganglionosis
(Hosoda et al., 1994;
McCallion et al., 2003;
Yamada et al., 2006)
Ednrbs-l/s-l

(aka “piebald lethal”)


Spontaneous gene deletion
ENCCs exhibit
delayed entry into
the developing
hindgut and do
not fully populate
the distal colon
despite the fact
that they migrate
longer during
development.
Aganglionosis of distal
colon; Decreased intensity
of AChE fibers in proximal
intestine; Alterations in
neuron types in
ganglionated regions of
colon in homozygotes;
Decreased density of
ganglia in heterozygotes.
(Cantrell et al., 2004;
Fujimoto, 1988;
Hosoda et al., 1994;
Webster, 1973)
(Puffenberger et al., 1994)
Ednrbtm1Ywa / tm1Ywa

(aka Ednrb −/−)

Targeted gene knockout
ENCCs exhibit
delayed entry into
the developing
hindgut.
Aganglionosis in distal
colon; hypoganglionosis in
regions of small intestine
(Cantrell et al., 2004;
Hosoda et al., 1994)
FoxD3



Transcription
Factor
FoxD3tm3Lby/tm3Lby

(aka Foxd3 −/−)

Floxed allele in combination
with either Wnt1cre or
Ednrb-iCre
Maintenance of
progenitors;
control of
progenitor
proliferation,
neural patterning,
and glial
differentiation
Lack of neurons in the
entire gastrointestinal
track
(Mundell et al., 2012;
Teng et al., 2008)
Mutation not
reported in GI
patients
Gdnf



Glial cell
derived
neurotrophic
factor, Ligand
for Ret
Gdnftm1Lmgd / tm1Lmgd

(aka Gdnf −/−)


Targeted gene knockout
Gdnf promotes
survival,
proliferation,
differentiation
and migration.
Total intestinal
aganglionosis in
homozygotes;
hypoganglionosis
throughout intestine of
heterozygotes
(Gianino et al., 2003;
Pichel et al., 1996a,b;
Sanchez et al., 1996;
Shen et al., 2002;
Wang et al., 2010)
(Angrist et al., 1996;
Ivanchuk et al., 1996)
Gfra1


Gdnf family
receptor
alpha 1, co-
receptor for
Ret
Gfra1tm1Jmi / tm1Jmi

(aka Gfra1 −/−)


Targeted gene knockout
Mediates
signaling of Gdnf
through Ret.
Total intestinal
aganglionosis in
homozygotes.
Heterozygotes have
decreased neuron size but
normal numbers.
(Cacalano et al., 1998;
Enomoto et al., 1998;
Gianino et al., 2003;
Hansen and Li, 2012;
Wang et al., 2010)
(Eketjall and Ibanez, 2002)
Ihh

Indian hedge
Hog
Ihhtm1Amc/tm1Amc

(aka Ihh −/−)


Targeted gene knockout
Promotes
survival of a
subpopulation of
ENCCs.
Perinatal lethality in
homozygotes. Colonic
aganglionosis
(Ramalho-Santos et al., 2000) Not reported
in GI patients
Itgbl

Integrinβ1
subunit
Itgb1tm3Ref/tm1Ref

Tg HtPA-Cre


(aka Itgb1 BGeo/flox ;
Tg.HtPA-Cre)


Conditional allele used in
combination with Tg HtPA-
Cre removed exons 2–7 in
neural crest
ENCCs do not
migrate beyond
mid hindgut.
ENCCs exhibit
altered migration
and increased
aggregation.
Defect in colonization of
cecum and proximal
hindgut; Altered ganglion
geometry.
(Breau et al., 2009;
Breau et al., 2006)
Not reported
in GI patients
Pax3Sp/Sp

Paired Box 3,
Transcription
Factor
Pax3Sp/Sp



Spontaneous Null allele
ENCCs fail to
enter fetal
intestine.
Total intestinal
aganglionosis in
homozygotes;
(Lang et al., 2000;
Lang and Epstein, 2003)
Mutation not
reported in GI
patients
Pds5b

Cohesion
regulatory
protein
involved in
sister
chromatid
cohesion
Pds5btm1Jmi/tm1Jmi

(aka Pds5b −/−)


Targeted gene knockout
Delayed entry of
ENCCs into distal
hindgut by 12dpc;
Alterend neuronal
density in regions
of small intestina
that are colonized
Perinatal lethality with
colonic aganglionosis.
(Zhang et al., 2007) Mutations of
PDS5B in
Cornelia de
Lange
patients do
not exhibit
aganglionosis.

Phox2b



Paired-like
homeobox 2,
Transcription
Factor
Phox2btm1Jbr / tm1Jbr

(akaPhox2b−/−)



Targeted gene knockout

Phox2btm2Heno/+ and
Phox2btm1Heno/+


(aka Phox2b del5/+
and Phox2b del8/+)


931 del5 and 693–700 del8
mutations lacking
nucleotides 931–935 and
693–700 in the ORF
Promotes
survival of
autonomic
progenitors.




Impaired
neuronal
differentiation
and decrease
proliferation of
the enteric
ganglion
progenitors
Total intestinal
aganglionosis in
homozygotes;





Perinatal lethality of
heterozygous
mutants;.Hypoganglionosis
in Phox2bdel5/+ and
colonic aganglionosis in
Phox2bdel8/+ and general
hypoganlionosis in other
gut segments
(Pattyn et al., 1999)





(Nagashimad a et al., 2012)
(Amiel et al., 2003)
Ret

Rearranged
during
transfection;
Receptor
Tyrosine
Kinase,
Rettm1Cos/tm1Cos

(aka Ret −/−)


Targeted gene knockout
Altered migration,
proliferation,
survival and
neuronal
differentiation of
ENCCs .
Total intestinal
aganglionosis in
homozygotes; In
heterozygotes normal
neuronal numbers,
decreased neuron size
and altered cholinergic
fibers are reported.
(de Graaff et al., 2001;
Gianino et al., 2003;
Schuchardt et al., 1994)
(Attie et al., 1995;
Attie et al., 1994;
Edery et al., 1994;
Romeo et al., 1994)
Ret – hypomorphic
isoforms


Humanized monomeric
isoform, multiple alleles that
produce only one isoform
from the gene or that alter
phosphorylation of the
receptor including:
Decreased
ENCCs and
failure to migrate
into fetal hindgut;
compromised
neuronal survival.
Effects depend
on isoform that is
mutated
Colonic aganglionosis (de Graaff et al., 2001;
Schuchardt et al., 1994;
Uesaka et al., 2008)
Ret mi51/mi51, Ret9/−; Ret S697A;
Rettm3(RET)Jmi/+
RetDN/+



Targeted mutation of
cytoplasmic domain

uncharacterized

Total intestinal
aganglionosis in
homozygotes;
Heterozygotes severe
hypoganglionosis and
reduced nerve fiber
density in ganglionated
intestine

(Jain et al., 2004)
Rettm1Cti/+

(aka RetC620R/+)


Targeted amino acid
substitution
uncharacterized Total intestinal
aganglionosis in
homozygotes;
Heterozytoes exhibit
decreased neuron
numbers and fiber density
(Carniti et al., 2006)
Rettm3Cos/tm3Cos

(aka Ret S697A/S697A)


Targeted nucleotide
substitution
Delayed
migration of
ENCCs into
colon.
Homozygous mutants
exhibit aganglionosis in
mid and distal colon.
(Asai et al., 2006)
Sox10

Sry box 10,
Transcription
Factor
Sox10Dom/+

(aka “Dominant
Megacolon”)


Spontaneous mutation,
dominant negative
Decreased
ENCCs, Delayed
migration in
heterozygotes,
vagal cell death
in homozygotes.
Variable aganglionosis in
colon of heterozygotes
and imbalance of neuron
subtypes in proximal
ganglionated intestine.
Homozygotes have
complete intestinal
aganglionosis
(Kapur, 1999;
Musser et al., 2015;
Southard-Smith et al., 1998;
Walters et al., 2010)
(Pingault et al., 1998)

Seen in
patients with
Waardenburg
-Shah (WS4)
and PCWH
Syndromes
Sox10tm1Weg/+



(aka Sox10LacZ/+)


Targeted gene knockout,
haploinsufficient
Altered
specification of
neural crest
lineages;
Decreased
numbers of
ENCCs; Loss of
mutipotency in
ENCCs; Delayed
migration in
heterozygotes,
Aganglionosis in distal
colon of heterozygotes.
(Britsch et al., 2001;
Paratore et al., 2002;
Paratore et al., 2001)
TashT Tg(SRY-
YFP,Tyr)TashTNpln

(aka “TashT”)


Untargeted transgene
insertion in gene desert
near Fam162b leads to
overexpression
Slower migration
of vagal ENCCs
into the fetal
intestine. Normal
directionality,
proliferation, and
differentiation of
vagal ENCCs
was observed.
Aganglionosis in distal
colon of subsets of
homozygous transgenic
mice with higher
predominance in male
TashT mutants.
(Bergeron et al., 2015) Mutation not
reported in GI
patients.
Zeb2


Zinc finger E-
box binding
homeodomai
n 2;
Transcription
factor,
Zeb2 tm1.1Yhi/tm1.1Yhi,
Tg.Wnt1-Cre


(aka
Zeb2flox(ex7),Wnt1Cre
)


Targeted gene knockout; or
neural crest ablation of
floxed allele in combination
with Wnt1-Cre
Vagal neural
crest cells do not
develop or
delaminate from
the neural tube.
Zeb2 homozygotes are
lethal at E9.5;

Zeb2fl/(ex7)fl(ex7) colonic and
partial small intestinal
aganglionosis
(Van de Putte et al., 2007;
Van de Putte et al., 2003;
Zimmer and Puri, 2015)

Van de Putte, 2003;
Van de Putte, 2007
(Wakamatsu et al., 2001)

Phenotypes of spontaneous or engineered alleles in mice that lead to alterations ENS structure and the roles of each molecule in ENS development if known are listed.

**

Mutant strains are identified by Mouse Genome Nomenclature (http://informatics.jax.org) that designates the specific allele studied. Fields marked as "unknown" indicate data not reported, or mouse model not tested.

*

Only the first report of disease gene mutation identified in patients is listed due to space constraints.

Table 2.

Gene Alterations in Mice that Impact the ENS Without Overt Intestinal Aganglionosis

Gene Allele type** Role in ENS
development
Effect on Mature ENS References
Araf

Araf proto-oncogene
serine threonine
kinase
Araftm1Mmc / tm1Mmc/

(aka Araf −/−)

Targeted gene
knockout
Not described Enteric ganglia present but
functionally defective with
abnormal architecture.
Homozygous mutants
exhibit colonic distension
(Pritchard et al., 1996)
Ascl1

achaete-scute family
bHLH transcription
factor
Ascl1tm1And / tm1And

(aka Mash1−/−)

Targeted gene
knockout
Promotes survival and
development of neuronal
subtypes
Decreased neuronal
numbers; wide spacing and
erratic arrangement of
enteric ganglia
(Blaugrund et al., 1996)
Bmp4

Bone morphogenetic
protein 4
Bmp4

Tg.(Eno2-
Bmp4)3Jake

transgene over-
expression in neurons
and entero-endocrine
cells using Neuron
Specific Enolase
promoter
Not described Regional specific increases
in 5HT+, dopaminergic, and
TrkC+ neurons
accompanied by increased
glial numbers
(Chalazonitis et al., 2011a;
Chalazonitis et al., 2008)
Celsr3

Cadherin, EGF LAG
seven-pass G-type
receptor 3
Celsr3tm1Agof/tm1Agof

(aka Celsr3 −/−)

Targeted gene
knockout
Disrupted neurite
network with fewer and
shorter longitudinal
processes and more
circumferential and
orally projecting
processes on
developing enteric
neurons. Normal
differentiation and
proliferation of neurons.
Perinatal lethal. Defects in
longitudinal tract formation
and neurite organization at
P0.
(Sasselli et al., 2013)
Celsr3
tm2Agof/tm2Agof,
Tg.Wnt1-Cre)

(aka Celsr3 flox/flox,
Wnt1Cre)

Floxed allele used in
combination with
Wnt1-Cre
Not described Required for organization of
enteric plexus.
Interganglionic strands are
reduced in thickness and
have irregular trajectories.
Aberrant Colonic Migrating
Motor Complexes and
deficient ability to propel
luminal contents in the distal
bowel.
(Sasselli et al., 2013)
Cdh2

N-Cadherin cell
adhesion molecule
Cdh2 tm1Glr / tm1Glr ;
Tg.Ht-PA Cre

(aka Ncad flox/flox)

Conditional floxed
allele used in
combination with Ht-
PA Cre
Promotes migration Delayed but complete gut
colonization
(Broders-Bondon et al., 2012)
Dicer1

Ribonuclease type III
processes miRNAs
Dicer1tm1Bdh / tm1Bdh
; Tg.Wnt1-Cre

(aka Dicerflox/flox;
Tg.Wnt1-Cre)

Conditional inactivation
of Dicer with Wnt1-Cre
Controls differentiation
and cell survival.
ENCCs migrate
normally, but marked
reductions in cell
numbers is noticed by
E17 with increased
apotosis.
Homozygous lethal at birth. (Zehir et al., 2010)
Dlx2

distal-less homeobox
2, Homeobox
transcription factor
Dlx2

(aka Dlx2 −/−)

Targeted gene
knockout
Not described Enteric neurons are present
but newborn pups die
perinatally with massive
distension of the intestinal
lumen
(Qiu et al., 1995)
ErbB2

Receptor of
neuregulin1
ErbB2tm1KLee/
tm1Klee; Tg.Nestin-
Cre

(aka ErbB2 −/−)

conditional floxed
allele used in
combination with
Nestin-Cre
Promotes postnatal
survival of neurons and
glial cells
Postnatal loss of neurons
and glial cells in the colon
upon deletion of ErbB2 in
epithelial cells
(Crone et al., 2003)
ErbB3

erb-b2 receptor
tyrosine kinase 3;
ErbB3tm2Cbm /
tm2Cbm

(aka ErbB3 −/−)
Aberrant enteric glial
development
Absent enteric glia (Chalazonitis et al., 2011b;
Riethmacher et al., 1997)
Fgf2

Fibroblast growth
Factor 2
Fgf2tm1Zllr / tm1Zllr

(aka Fgf2 −/−)

Targeted gene
knockout
Unknown Hyperplastic enteric ganglia
with open architechture of
connectives ; altered
mucosal barrier function and
Cl- secretion.
(Hagl et al., 2008;
Hagl et al., 2013)
Fzd3

frizzled class receptor
3 , Wnt receptor
F zd3tm1Nat/tm1Nat

(aka Fzd3 −/−)
Required for guidance
and growth of enteric
neuronal projections.
Normal distribution of
TuJ1+ neurons
compared to controls,
however the neurite
network is disrupted.
Defects in longitudinal tract
formation and neurite
organization at P0
(Sasselli et al., 2013)
Gas1

Growth arrest
specific1; blocks
entry to S phase and
plays a role in growth
suppression
Gas1tm2Fan/ tm2Fan


(aka Gas1LacZ /LacZ
)

Targeted gene knock-
in
Repels enteric axons Homozygous lethal at birth;
Increased numbers of
enteric neurons,
mislocalized within gut
mesenchyme;
(Biau et al., 2013;
Jin et al., 2015)
Gdnf

glial cell line derived
neurotrophic factor
Gdnf

Tg.(Myog-
Gdnf)1Lich

Gdnf transgene over
expression in muscle
Not described Increased numbers
submucosal neurons;
changes in fiber density;
accelerated intestinal transit
(Wang et al., 2010)
Gfra2

Gdnf family receptor
alpha 2
Gfra2tm1Msa/ tm1MSa

(aka Gfra1 −/−)

Targeted gene
knockout
Promotes neuron
survival and outgrowth
of neurites
Reduction of fibers and
abnormal motility, defective
pancreas innervation
(Rossi et al., 2003;
Rossi et al., 1999)
Hand2

Heart and neural
crest derivatives
expressed transcript
2; Transcription factor
Hand2
tm1Majh/tm1Majhor

Hand2tm1Cse/tm1Cse

(aka Hand2 −/−;
Tg.Wnt1-Cre)

conditional floxed
allele crossed to Wnt1-
Cre
Promotes terminal
differentiation of VIP and
nNOS neuronal
subtypes
Disorganized ENS plexus,
decreased neuronal
numbers, reduced GI transit
(D’Autreaux et al., 2011;
Hendershot et al., 2007;
Lei and Howard, 2011;
Morikawa et al., 2007)
Hipk2

Homeodomain
interacting protein
kinase 2
Hipk2tm1Ejh/tm1Ejh

(aka Hipk2 −/−)
Not described Postnatal increase in glia
and progressive loss of
neurons with concurrent
arrest of synaptic maturation
in enteric neurons
(Chalazonitis et al., 2011b)
Hlx1

H2.0-like homeobox,
Transcription factor,
Hlx1tm1Rph/ tm1Rph

(aka Hlx1 −/−)

Targeted gene
knockout
Reduction of ENCCs in
fetal foregut, severe
reduction of ENCCs in
mid- and hindgut.
Homozygotes embryonic
lethal with severe
hypoganglionosis
(Bates et al., 2006)
Hoxb5

Homeobox B5 –
engrailed fusion,
transcription factor
Tg.(CAG-
en/Hoxb5-
EGFP)#Vchl

dominant negative
allele; inducible by Cre
Dominant negative allele
antagonizes normal
Hoxb5; Reduction of Ret
expression, reduced
migration of ENCCs
Reduction of neurons or
aganglionosis affecting
colon and ileum
(Lui et al., 2008)
Kif26a

Kinesin family
member 26 a; motor
protein
Kif26atm1.1Noh/
tm1.1Noh

(aka Kif26a −/−)

Targeted gene
knockout
Negative regulator of
RET signaling
Enteric neuronal hyperplasia
with pseudoobstruction
(Zhou et al., 2009)
L1cam

L1 cell adhesion
molecule
L1camtm1Mtei /
tm1Mtei

(aka L1cam −/−)
Targeted gene
knockout
Enables ENCC
migration
Delayed but complete gut
colonization.
(Anderson et al., 2006)
Noggin

noggin protein
Noggin

Tg.(Eno2-Nog-
EGFP)Alch

(aka NSE-
Noggin)

Noggin transgene over
expression in neurons
and enteroendocrine
cells using Neuron
Specific Enolase
promoter
Promotes development
of TrkC+ neurons
Increased neuronal
numbers in myenteric and
submucosal plexi ; reduction
of TrkC+ neurons, altered
proportions of 5HT+,
GABA+ and CGRP+
neurons ; altered intestinal
transit and stool composition
(Chalazonitis et al., 2004;
Chalazonitis et al., 2011a;
Chalazonitis et al., 2008)
Nrtn

Neurturin
Nrtntm1JMi / tm1Jmi

(aka Nrtn −/−)

Targeted gene
knockout
Ret Ligand; Promotes
neurite outgrowth
Homozygotes have reduced
neuron soma size, neuron
fiber density, abnormal GI
motility
(Heuckeroth et al., 1999;
Yan et al., 2004)
Ntf3

Neurotrophin-3,
ligand for p75 low
affinity neurotrophin
receptor
Ntf3tm1Par / tm1Par

(aka Ntf3 −/−)

Targeted gene
knockout
Expressed by
mesenchymal cells;
promotes survival and
differentiation of
developing neurons
Region specific decrease in
enteric neuron number
(Chalazonitis et al., 2001)
Ntrk3

Receptor of
neurotrophin-3
Ntrk3tm1Par / tm1Par

(aka TrkC −/−)

Targeted gene
knockout
unknown Reduction of myenteric and
submucosal neuron
subtypes
(Chalazonitis et al., 2001)
Phactr4

Phosphatase and
actin regulator 4
Phactr4humdy/humdy

ENU mutation
missense change in C-
terminus; hypomorphic
allele
Promote migration;
regulates directional
migration of ENCCs
through PP1,
Itgb1signaling; reduced
number of NADPH-
diaphorase stained
neurons in the E18
colon
Homozygous lethal at birth (Zhang et al., 2012;
Zhang and Niswander, 2012)
Pofut1

Protein O-
fucosyltransferase 1
Pofut1tm1Ysa / tm1Ysa
;Tg.Wnt1-Cre

(aka Pofut1 flox/flox
)

Conditional Floxed
allele crossed to Wnt1-
Cre
Promotes proliferation
and glial development
Premature neurogenesis
and reduction of glial cells;
hypoganglionosis
(Okamura and Saga, 2008)
Ptch1

Patched 1
Ptch1tm1Bjw / tm1Bjw

(aka Ptch1 −/−)

targeted gene
knockout
Reduced proliferation of
ENCCs; Reduced
neurogenesis;
Premature gliogenesis
Embryonic lethal at 12.5dpc (Ngan et al., 2011)
Pten

Phosphatase and
Tensin homology
Ptentm1Hwu / tm1Hwu

(aka Pten flox/flox;
Tg. Tyr-Cre)

conditional inactivation
using Tyr-Cre
Inhibits migration and
proliferation
Hypertrophy and
hyperplasia of the ENS; fatal
intestinal pseudo-
obstruction
(Puig et al., 2009)
Rb1

Retinoblastoma
protein1
Rb1tm2Brn / tm2Brn,
Tg(Tyr-Cre) (aka
Rb1flox/flox, Tg(Tyr-
Cre))

Conditional inactivation
using Tyr-Cre
Not described Abnormal intestinal motility;
increased enteric glia;
nNOS neurons exhibit large
nuclei as a result of DNA
replication without cell
division
(Fu et al., 2013)
Shh

Sonic Hedghog
Shhtm1Amc / tm1Amc

(aka Shh −/−)

targeted gene
knockout
Expressed by epithelial
cells; promotes
proliferation and
concentric patterning
Increased number of
neurons in mucosa of
knockout mutant mice
(Ramalho-Santos et al., 2000)
Slc6a2

solute carrier family 6
(neurotransmitter
transporter,
noradrenalin),
member 2
Slc6a2 tm1Mca
/tm1Mca

(aka Net −/−)

Targeted gene
knockout
Unknown Decreased numbers of
myenteric neurons ;
Reduced numbers of
neuronal subsets including
5HT+ and calretinin+
(Li et al., 2010)
Spry2

Sprouty homolog 2
Spry2tm1Ayos/tm1Ayos

(aka Sprouty2 −/−)

Targeted gene
knockout
Negative regulator of
Gdnf signaling; Alters
neonatal development or
survival of enteric
neurons
Enteric neuron hyperplasia,
oesophagal achalasia
(Taketomi et al., 2005)
Sufu

Suppressor of Fused

negative regulator of
Gli transcription
factors
Sufu tm1Hui / tm1Hui;
Tg.Wnt1-Cre

Sufu flox/flox;Wnt1-
Cre

Conditional invalidation
using Wnt1-Cre
Neurons and glia highly
disorganized in all Sufu
mutants at E13.5;
neuron to glia ratios
reduced. Severe axonal
fasciculation defects;
delay in gut colonization
observed in 14% of
mutants
Unknown- Sufu mutants
died

around E14 but show
hypoganglionosis
(Liu et al., 2015)
Tcof1

Treacher Collins
Franceschetti
syndrome1, homolog;
nucleolar factor
Tcof1 tm1Mjd/+ (aka Tcof
−/+)

Targeted gene
knockout
Delayed and prolonged
migration of ENCCs
Neonatal lethality of
heterozygotes prevents
functional analysis of any
motility deficits.
(Barlow et al., 2012)
Tfam

Transcription factor
A, mitochondrial
Tfamtm1Lrsn/tm1Lrsn
;Tg.CNP-Cre

(aka Tafm flox/flox)

Floxed allele crossed
to CNP-Cre
Not described Postnatal death of specific
subsets of enteric neurons
(Viader et al., 2011)
Tlr2

Toll-like receptor 2
Tlr2tm1Kir / tm1Kir

(aka Tlr2 −/−)

Targeted gene
knockout
Not described Aberrant connectivity of
enteric ganglia; Altered
intestinal contractility;
Abnormal mucosal
secretion; Altered
inflammatory responses in
colon model inflammatory
bowel disease
(Brun et al., 2013)
Tlx2

T cell leukemia,
homeobox 2, Hox
Transcription factor
Tlx2tm1Sjk / tm1Sjk
(aka Hox11L1 −/−)

Targeted gene
knockout
Not described Hyperinnervation of the
proximal colon accompanied
by megacolon at 3–5 weeks
of age
(Hatano et al., 1997;
Shirasawa et al., 1997)
Tph2

Tryptophan
Hydroxylase 2,
neuronal expressed
form of enzyme
critical for serotonin
synthesis
Tph2tm1Lex/tm1Lex

(aka Tph2 −/−)

Targeted gene
knockout
Promotes differentiation Homozygotes show
decreased neuronal density
in the ileum, reduced
proportions of dopaminergic
neurons with resulting
slowed GI transit and
colonic emptying.
(Li et al., 2011)
Zic2

Zinc finger protein of
the cerebellum 2,
transcription factor
Zic2m1Nisw/m1Nisw

ENU-induced
missense mutation
Potential negative
regulator of nerve fiber
growth.
Homozygous lethal.
Increased density of enteric
ganglia and neurite
extension
(Zhang and Niswander, 2013)
**

Mutant strains are identified by Mouse Genome Nomenclature (http://informatics.jax.org) that designates the specific allele studied. Fields marked as "unknown" indicate data not reported, or mouse model not tested.

Table 3.

Environmental Factors that Impact the ENS in Mice

Environmental
factor
Gene Allele
type**
Role in ENS
development
Effect on Mature
ENS
References

Vitamin A
Aldh1a2

Aldehyde
dehydrogenase
family 1, enzyme
essential for
production of
retinoic acid
Aldh1a2
tm1lpc /tm1lpc

(aka
Raldh2−/−)

Targeted
gene
knockout
Severe reduction of
post-otic vagal neural
crest and absence of
Ret+ ENCCs in foregut
Homozygous lethal (Niederreither et al., 2003)
MPA
mycophenolic
acid
Impdh2

inosine 5’
monophosphate
dehydrogenase
Impdh2
tm2Bmi /
tm2Bmi;
Tg.Wnt1-
Cre

(aka
Impdh2
flox/flox,
Wnt1-Cre)
Reduces the number of
migrating ENCC with
lamellipodia, DNA
synthesis and induces
ENCC apoptosis.
Mycophenolate
treatment induces
bowel aganglionosis
and increases the
penetrance and
severity of HSCR of
Ret9 and Sox10lacZ
mutants

Impdh2 mutants
present with multiple
NCC defects,
including highly
penetrant intestinal
aganglionosis,
(Lake et al., 2016;
Lake et al., 2013)
Vitamin A Rbp4

Retinal binding
protein 4 plasma;
binds Vitamin A in
serum
Rbp4
tm1Gott /
tm1Gott

Targeted
gene
knockout

(aka Rbp4
−/−)
In absence of dietary
Vitamin A, mutants lack
ENCC migration into
hindgut; Vitamin A
affects lamellipodia
formation and ENCC
migration in response
to GDNF
Colorectal
aganglionosis; distal
from stomach if
deprived of dietary
retinoid from E7.5
(Fu et al., 2010)
Ibuprofen Gene unknown Reduced migration in
Ret +/− heterozygotes
with decreased
lamellipodia and levels
of RAC1/CDC42.
(Schill et al., 2016)

Table 4.

Summary of Genetic Interactions That Impact the ENS Based on Double Mutant Mouse Studies.

Interacting
Genes
Alleles used Effect on mature ENS and
developmental origin
Molecular
basis
References
Ret × Ednrb Ret +/− Ednrbs,
Ednrbsl
Highly penetrant aganglionosis present in
double mutants.
unknown (Carrasquillo et al., 2002;
McCallion et al., 2003)
Ret × Edn3 Ret51 Edn3ls Removal of Edn3 activity in Ret51/51 animals
results in increased intestinal aganglionosis
length compared to Edn3-deficient or Ret51/51
embryos.
Presence of a single Ret51 allele partially
rescues the aganglionosis of Edn3ls/ls
animals.
Synergistic effect and antagonistic roles of
these pathways on proliferation and migration
of ENS progenitors, respectively.
Intercrossed
signaling pathways-
PKA involvement
(Barlow et al., 2003)
L1cam ×
Ednrb
L1cam −/y Ednrbs,
Ednrbsl,
Loss of L1cam exacerbates the
hypoganglionosis in the caudal colon of
Ednrbs/s mice.
Haploinsufficiency of L1cam increase the
severity of the aganglionosis in Ednrbsl/sl.
unknown (Wallace et al., 2011)
L1cam ×
Edn3
L1cam −/y Edn3+/− Loss or haploinsufficiency of L1cam
increases the severity of the aganglionosis in
Edn3 mutant mice. Double mutants also
present with hypoganglionosis.
unknown (Wallace et al., 2011)
Sox10 ×
Ednrb
Sox10LacZ
Sox10Dom
Ednrbs,
Ednrbsl,
Double heterozygotes present with increased
intestinal aganglionosis length compared to
Sox10 heterozygotes.
Defects partly due to increased cell death of
vagal crest cells prior to entry into foregut and
increased neuronal differentiation.
SOX10 activates
Ednrb enhancer in
vivo
(Cantrell et al., 2004;
Stanchina et al., 2006;
Zhu et al., 2004)
Sox10 ×
Edn3
Sox10LacZ
Sox10Dom
Edn3ls Double heterozygotes and double mutants
present with increased intestinal
aganglionosis length compared to single
mutants.
SOX10 activates
Ednrb enhancer
(Stanchina et al., 2006;
Zhu et al., 2004)
Sox10 ×
Sox8
Sox10LacZ Sox8LacZ Loss or haploinsufficiency of Sox8 increases
the severity of the aganglionosis in Sox10
heterozygotes.
Defects are partly due to increased cell death
of vagal crest cells before gut entry and
reduction of glial cells in double
heterozygotes.
unknown (Maka et al., 2005)
Sox10 ×
Zeb2
Sox10LacZ Zfhx1bΔex7 Zeb2 heterozygosity increases the severity of
the aganglionosis in Sox10 heterozygotes
from E11.5 onwards. Double heterozygotes
also present with hypoganglionosis.
Defects are partly due to decreased
progenitor proliferation and increased
neuronal differentiation.
unknown (Stanchina et al., 2010)
Sox10 ×
L1cam
Sox10LacZ L1cam−/y Loss of L1cam increases the penetrance of
the aganglionic phenotype observed in
Sox10+/lacZ embryos from E11.5 onwards.
Double mutants present with additional
hypoganglionosis.
Defects partly due to increased cell death of
vagal crest cells prior to entry into foregut.
SOX10 regulates
L1cam expression
(Wallace et al., 2011)
Sox10 ×
Itgb1
Sox10LacZ
Sox10Dom
Ht-PA-Cre;
Itgb1f/neo
Double mutants present with increased
intestinal aganglionosis length and more
severe neuronal network disorganization from
E11.5 onwards, due to altered cell migration
capacities
Unknown (Watanabe et al., 2013)
Sox10 ×
Ets1
Sox10LacZ Variable
spotting
Hypoganglionosis in Sox10LacZ/+ ; Ets1−/− Est1 activates Sox10
enhancer element
MCS4/U3 in vitro
(Saldana-Caboverde et al., 2015)
Sox10 ×
Sufu
Sox10NGFP Wnt1-Cre;
Sufuflox / flox
Reduction of Sox10 attenuates glial
differentiation defects of Sufu mutants.
Gli1 and Gli2 activate
SOX10 MCS4/U3
and MCS7/U1
enhancers + SOX10
represses Sufu
(Liu et al., 2015)
Tcof × Pax3 Tcof1+/− Pax3Sp/+ Colonic aganglionosis in double
heterozygotes.
Phenotype partly due to cumulative apoptosis
and decreased proliferative capacity.
unknown (Barlow et al., 2013)
Cdx × Pax3 Gt(ROSA)26
Sortm1(en/Cdx1−
EGFP)Npln;
Pax3Pro-Cre

(aka,
R26REnRCdx1,
P3Pro-Cre)

Dominant
negative
Engrailed
knockdown of
Cdx1 in
neurectoderm
Pax3Sp/+ Hypoganglionosis is present in double
heterozygotes.
Transactivation of a
neural crest
enhancer of Pax3 by
Cdx
(Sanchez-Ferras et al., 2016;
Sanchez-Ferras et al., 2012)
Itgb1 × Cdh2 Ncadfl/fl Ht-PA∷cre;
Itgb1 f/neo
Double mutants present with increased
intestinal aganglionosis length, but rescued
Itgb1 aggregation anomalies and neuronal
network disorganization.
Phenotype partly due to altered speed of
locomotion and directionality as well as
changes in the balance of adhesion.
unknown (Broders-Bondon et al., 2012)

Crosses performed and effects on mature ENS as well as developmental and molecular origin of defects observed are listed. Fields marked as “unknown” indicate data not reported.

The first mouse model of HSCR was generated by targeting the Ret gene (Schuchardt et al., 1994). This tyrosine kinase receptor interacts with four distinct ligands [glial cell line-derived neurotrophic factor (Gdnf), neurturin (Nrtn), artemin (Artn) and persephin (Pspn)]. Each of these activates Ret by binding to the glycosylphosphatidylinositol-linked Gdnf family of co-receptors (Gfra1 to 4). In mice, total Ret deficiency causes complete intestinal aganglionosis. Ret−/− mice additionally present with kidney agenesis and die at birth. Gdnf and Gfra1 deletions cause nearly identical phenotypes, indicating that they are the critical Ret activators during fetal development (Table 1; (Cacalano et al., 1998; Durbec et al., 1996; Enomoto et al., 1998; Moore et al., 1996; Pichel et al., 1996a; Sanchez et al., 1996)). Gdnf haploinsufficiency also leads to severe hypoganglionosis (Table 1; (Gianino et al., 2003; Shen et al., 2002)). In contrast, mutants affecting other Ret ligands or co-receptors present with subtler defects, including reduced nerve fiber density, abnormalities in neurotransmitter release, or hypoganglionosis (Table 2, and see for example (Obermayr et al., 2013; Young, 2012; Zimmer and Puri, 2015)). A large variety of Ret mutant mice have been generated over time. These new alleles include mono-isoformic variants as well as serine and tyrosine phosphorylation mutation sites (Table 1). Each of the latter, as well as inactivation of Ret inhibitors, greatly helped decipher downstream signaling pathways involved in ENS ontogenesis (Table 1 and 2 and references therein). In total, data from these models show that Ret signaling is essential for ENS precursor proliferation, migration, differentiation, survival, and neurite growth. Gdnf/Ret signaling can also influence specific subtypes of neurons, with reduction of neuronal nitric oxide synthase (nNOS) in some mutants (Roberts et al., 2008; Uesaka and Enomoto, 2010). Interestingly, conditional inactivation of Ret or Gfrα1 after gut colonization by ENCCs causes loss of neurons in the colon, suggesting that this signaling pathway is also essential for survival of colonic ENCCs (Uesaka et al., 2007; Uesaka et al., 2008).

Other cell populations in addition to vagal and sacral neural crest-derived progenitors contribute to formation of enteric ganglia and are Ret signaling dependent. Using genetic fate mapping in mice, Enomoto’s group indeed definitively demonstrated that a subset of Schwann cell precursors (SCPs), that invade the gut along extrinsic nerves, adopt a neuronal fate in the postnatal period and contribute to the ENS (Uesaka et al., 2015; Uesaka et al., 2016). Genetic ablation of Ret in SCPs caused colonic aganglionosis, indicating that SCP-derived neurogenesis is essential for ENS integrity, providing novel insight into the development and disorders of neural crest-derived tissues.

Endothelin-3 (Edn3, a member of the 21 amino acid family of peptides, processed by the Ece1 enzyme) and its seven transmembrane G-coupled receptor (Ednrb) are members of a second pathway shown to play crucial roles in ENS development (Table 1, and references therein). Point mutations in Edn3 and/or deletion of Ednrb are causative for spontaneous mouse mutants called Lethal spotting, Piebald lethal and Piebald respectively (Baynash et al., 1994; Hosoda et al., 1994). The first two mutants, in a manner similar to the genomic knock-out of these genes, present with delay in ENCCs migration within the small intestine, and distal hindgut aganglionosis (Barlow et al., 2003; Baynash et al., 1994; Hosoda et al., 1994). All mutants in the endothelin pathway additionally present with pigmentation defects due to abnormal development of neural crest-derived melanocytes. Consistent with the phenotypes observed in mouse models, mutations in either gene have been found causative for isolated HSCR or Waardenburg-Hirschsprung disease (HSCR combined with pigmentation defects and deafness) in humans (Table 1, and for reviews see (Amiel et al., 2008; Pingault et al., 2010). During ENS development, the primary function of Ednrb and its ligand is to prevent neuronal differentiation of ENCC progenitors and to keep them in a proliferative state, thereby maintaining a pool of uncommitted cells with potential to colonize the gut (Bondurand et al., 2006; Hearn et al., 1998; Nagy and Goldstein, 2006; Wu et al., 1999). Endothelins also appear to promote ENCCs migration directly and perturbation of Ednrb principally affects individual ENCC speed rather than directionality (Druckenbrod and Epstein, 2009; Young et al., 2014), however the underlying mechanism is still not clear.

Over the years, a cohort of transcription factors, including Sox10, Phox2b, Zeb2, as well as Foxd3, Hand2, Ascl1 (formerly known as Mash1), Pax3, Tlx2 (formerly Hox11L1), Hoxb5, Hlx1, Dlx2 and interacting cofactors such as HipK2 have been implicated in ENS development. Although each of the corresponding mouse models present with severe ENS defects (see Table 1 and 2 and references therein), only mutations within the first 3 genes lead to aganglionosis and are observed in syndromic forms of HSCR in humans, namely Waardenburg-Hirschsprung disease, Haddad Syndrome (Central Congenital Hypoventilation Syndrome associated with HSCR) and Mowat-Wilson Syndrome (Intellectual disabilities associated with specific craniofacial abnormalities and HSCR in about 40% of cases), respectively. Homozygous loss of Phox2b in mice leads to a complete absence of ENS similar to the phenotype of the Ret mutants (Pattyn et al., 1999). More recently, generation of mutants bearing variations identical to the ones identified in CCHS patients (931 del5 and 693–700 del8), revealed heterozygous animals failing to breath and presenting with colonic aganglionosis or hypoganglionosis (Nagashimada et al., 2012). The mutations demonstrate that Phox2b is essential for normal ENS development but also showed Phox2b variants with altered transcriptional activation protein domains can function as dominant negatives leading to severe deficits in ENS development with resulting aganglionosis. The development of the ENS is even more exquisitely sensitive to levels and function of Sox10 and Zeb2 (Table 1 and references therein). Haploinsufficiency of Sox10 leads to distal colonic aganglionosis, while homozygous mutants die between E13.5 and birth, with total absence of ENCCs even within the esophagus (Herbarth et al., 1998; Kapur, 1999; Southard-Smith et al., 1998). Sox10 is a member of the Sry related family of transcription factors that is expressed early in neural crest development, including in the vagal and sacral regions, and later in ENCCs when they reach the gut and migrate along it. Similar to Edn3 signaling, Sox10 promotes the maintenance of a pool of progenitor cells for ENS colonization (Bondurand et al., 2006; Kim et al., 2003). Suppression of Sox10 expression is required for neuronal differentiation to occur and complete absence of ENCCs within the digestive tract of homozygous mutants is due to early cell death of vagal crest cells before they reach the gut (Kapur, 1999; Southard-Smith et al., 1998), illustrating the critical requirement of this transcription factor for ENCC survival and differentiation. More recently, Sox10 was also shown to control cell migration (Corpening et al., 2011) by modulating cell adherence properties (Watanabe et al., 2013). In vitro studies indicate that Sox10 in concert with Pax3 regulates expression of Ret, and additional molecular studies in vitro suggest that Sox10 is capable of binding regulatory regions for Ednrb and Sox10 itself (Lang et al., 2000; Wahlbuhl et al., 2012; Zhu et al., 2004). While transcription factors that participate with Sox10 auto-regulation have been identified (Wahlbuhl et al., 2012), cofactors that participate with Sox10 in regulation of Ednrb are not yet known. Analysis of Sox10Dom mouse mutants revealed that alterations in this transcription factor not only lead to distal aganglionosis, but can also disrupt normal proportions of neuron subtypes in ganglionated small intestine that lead to deficits in intestinal transit (Musser et al., 2015). Similarly, Zeb2 plays essential roles in enteric neurogenesis by promoting ENCC proliferation and early migration within the fetal bowel (Stanchina et al., 2010; Van de Putte et al., 2003) via interactions with Sox10.

In addition to transcription factor networks, cell surface molecules, including the cell adhesion molecule L1CAM, β1Integrins, and N-cadherin, also play crucial roles during ENS cell migration and adhesion processes (Tables 1 and 2 and references therein). The gut microenvironment expresses several extracellular matrix components (ECM) molecules including fibronectin, laminins, collagens, tenascin-C and proteoglycans of various families (Akbareian et al., 2013; Breau et al., 2009). The major ECM receptors are integrins, and in 2006 deletion of Itgb1, encoding the Integrinβ1 subunit, in ENCCs was performed (Breau et al., 2006). Abnormal cellular adhesion, delayed migration and distal aganglionosis was documented in the resulting mutants. The migratory defect seems to occur specifically in the cecum/proximal hindgut and is thought to be due to a requirement for Integrinβ1-mediated interactions between ENCC and extracellular matrix components tenascin-C and fibronectin (Breau et al., 2009). More recently, studies that analyzed N-cadherin and Integrinβ1 interactions highlighted the complex regulation that exists between cell-cell and cell-matrix adhesion molecules and revealed how the correct balance between these two types of adhesion processes is crucial for ENS ontogenesis (Broders-Bondon et al., 2012).

In 2015, a novel role for collagen, the most abundant extracellular matrix protein, has been identified in ENS development. Soret and colleagues identified a new mechanism that causes HSCR-like disease in mice and involves deposition of excess collagen VI in the intestine by migrating ENS precursors as they colonize the fetal bowel (Soret et al., 2015). The description of a new mouse model, named Holstein, which was generated through an insertional mutagenesis screen that led to altered gene expression of the collagen-6α4 (Col6a4) gene is described further in the section entitled “Beyond genes: role of regulatory sequences, miRNAs and environmental factors”.

Besides these well-known molecules, additional factors have been implicated in the development of the ENS from animal studies including Neurotrophin-3, Sonic hedgehog (Shh), Indian Hedgehog (Ihh), Bone morphogenic proteins (Bmp2 and 4), Notch, Small GTPases, Neuregulin (Nrg1), Serotonin, planar cell polarity genes Celsr3 and Fzd3, nucleolar protein Treacle (Tcof1) and Sprouty 2 among others (see Tables 1 and 2 for lists of genes and functional effects on the ENS). Some of these mice recapitulate simple aganglionosis that is the hallmark of HSCR (Table 1). However, a number of them, are models of more complex gastrointestinal (GI) disease, affecting only specific neuronal subtypes or glial cells in late embryogenesis or adulthood and/or leading to hypo- or hyperganglionosis (Table 2 and references therein). For example, inactivation of negative regulators of the Ret signaling pathway (such as Sprouty 2 or Kinesin superfamily protein 26A; Kif26A) lead to hyperganglionosis and motility disorders due to over activated Ret signaling, such as esophageal achalasia (Taketomi et al., 2005; Zhou et al., 2009). Use of conditional gene inactivation to specifically ablate some of these signaling pathways in ENCCs or in the developing fetal bowel when ENCCs are emigrating has provided greater support for their involvement in ENS development. This is the case for intracellular components of signaling pathways with wide spatio-temporal functionality, such as Retinoblastoma protein (Rb1), Rho GTPases, the miRNA processing enzyme Dicer, the phosphatase and actin regulator Pten, the Tryptophan hydroxylase 2 enzyme Tph2 (Fu et al., 2013; Fuchs et al., 2009; Huang et al., 2010; Li et al., 2011; Puig et al., 2009) and more recently Suppressor of Fused (Sufu) (Liu et al., 2015). The recent identification of missense mutations affecting the GLI1, 2 and 3 transcriptions factors in patients presenting with HSCR and the analysis of the Sufu mutants indeed confirmed the essential role of the whole hedgehog signaling pathway (Liu et al., 2015). Hedghog proteins have important roles as morphogens. Both ligands are expressed in the gut, but have very different effects. Targeted Shh mutation results in ectopic enteric neurons (Shh has been shown to promote proliferation, inhibits neuronal differentiation and prevent premature centripetal invasion of gut by ENCC), whereas loss of Ihh causes aganglionosis in parts of the gut (see Table 2 and references therein). Ablation of Gas1 that mediates Shh signaling exhibits ENS deficits comparable to targeted deletion of Shh (Biau et al., 2013). Ectopic expression of the transcriptional effector Gli1 produces an effect similar to loss of Ihh, with hypoganglionosis (Yang et al., 1997). This data well correlates with the five gain of function missense mutations newly identified in GLI1, 2 and 3 in HSCR patients (Liu et al., 2015). Although prevalence of these mutations is surprisingly high, these data prompted Liu et al to further analyze the role of this signaling pathway during ENS development, and focused on Sufu, which negatively regulates the activities of GLI proteins. Previous studies demonstrated that hedgehog signaling is required for the normal migration of enteric neural progenitors and glial cell generation through Sox10 regulation (Liu et al., 2015; Ngan et al., 2011). In Sufu mutants, the networks of enteric ganglia were disorganized and neuron:glial cell ratios were reduced, a defect reminiscent of that recently described in Sox10 and Raldh1, -2, and -3 mutants (Musser et al., 2015; Wright-Jin et al., 2013). Because mice lacking Sufu in neural crest cells die before the gut is completely colonized by ENCCs, the authors were however not able to determine whether loss of Sufu results in an absence of neurons in the distal bowel and thus a HSCR-like phenotype. Altogether, these studies emphasize the importance of conditional models in ENCC developmental studies. Use of Cre driver lines such as Wnt1-Cre, Ht-PA-Cre, Nestin-Cre or Tyr-Cre that differ in their temporal and cell type specific expression of Cre can produce gene deletions that affect distinct phases of neural crest and ENCC development. These conditional approaches are complementary and essential for understanding the complexity of genes that contribute to HSCR; however caution is needed in comparing phenotypes across different crosses if the Cre line used differs between studies.

Recent studies of the ENS in mouse models have also identified roles for specific molecules in enteric neuron neurotransmitter specification, axon growth and navigation, target selection, synapse formation and development of mature electrical properties in late development and in adult (see Tables 2 and 3 and references therein). The diversity of molecules and mechanisms that contribute to the entire process of ENS ontogeny is not surprising as ENS maturation continues after the gut has been initially colonized by ENCCs, with additional neuron types exiting the cell cycle even after birth (Pham et al., 1991). While an increasing number of genes that contribute to the etiology of HSCR have been identified, surprisingly these genes account for only a small proportion of known cases. Generation of new models, gene interaction studies, and environmental factors that affect ENS development is therefore of high importance. Recently, the pace of identifying new mutations that lead to aganglionosis in mice has declined. This may be due in part to the oligogenic inheritance of the disease that greatly complicates generation of relevant genetic models and the time required for thorough functional validation. Fortunately, candidate gene approaches have been replaced by more general unbiased strategies. Indeed, a number of the recently published models were generated via insertional mutation or N-Ethyl-N-Nitrosourea (ENU) screens, looking for new neural crest regulators (Bergeron et al., 2015). This is the case of of the TashT and Holstein mouse models that will be described in the next section, and of Zic2m1Nisw presenting with enteric hyperplasia and dysplasia (Zhang and Niswander, 2013). Analysis of mutant phenotypes emerging from these screens will hopefully lead to the identification of novel HSCR genes in the coming years.

Beyond genes: roles for regulatory sequences, miRNAs and environmental factors

Beyond HSCR gene identification, human genetic studies have highlighted the major importance of variations within regulatory sequences of known genes as causative factors in aganglionosis. The increasing number of non-coding mutations within or near the RET locus are a reminder that although these types of changes may have low penetrance, they are capable of acting synergistically with other mutations to affect a disease phenotype (see for examples (Brooks et al., 2005; Emison et al., 2005; Griseri et al., 2007; Sribudiani et al., 2011). Deletion or point mutations identified within regulatory sequences of other genes including SOX10 and DSCAM are consistent with significant effects of regulatory variants on HSCR susceptibility (Jannot et al., 2013; Lecerf et al., 2014). Non-coding regulatory alterations around known HSCR genes and genes that may affect the ENS therefore need to be carefully considered, as some of the missing heritability in children with HSCR might result from noncoding variants that alter gene expression. In mice, at least three models bearing altered regulatory sequences have been described that present with neural crest defects, including aganglionosis phenotype, and are consistent with an effect of regulatory sequences perturbing essential gene function in the ENS. All three models were generated via insertional mutation screens looking for new neural crest regulators. In each case traditional “knockout” approaches would never have led to the discovery of these new HSCR susceptibility loci because the insertions alter noncoding regions of the genome that were poorly characterized at the time of publication and the affected genes are overexpressed in two of these models.

The first model, Sox10Hry, was published in 2006, well before the full description of Sox10 enhancers (Antonellis et al., 2006). This mutant presents with distal intestinal aganglionosis and severe hypopigmentation due to a 16 kb deletion upstream of the Sox10 gene that helped define and validate the importance of distant Sox10 regulatory elements. Over the last two years, the description of two new mouse models further validated the importance of regulatory sequences by identifying the Fam162b and Col6a4 genes (Bergeron et al., 2015; Soret et al., 2015). The first new mouse model, a transgenic line named TashT, displays a partially penetrant aganglionic megacolon phenotype shown to result from delayed migration of ENCCs, concomitant with insensivity towards Gdnf and Edn3 . This insertional mutation lies in a gene desert containing multiple highly conserved elements that exhibit repressive activity in reporter assays. RNASeq and 3C assays revealed the insertion results, at least in part, in loss of repression of the uncharacterized Fam162b gene in ENCCs. Interestingly, the partially penetrant aganglionic megacolon phenotype mostly affects males and may through future studies explain the prominent gender bias of 4:1 in males compared to females that is known to occur in human HSCR disease (for review see (Amiel et al., 2008)). The second model named Holstein, which was generated in the same genetic screen, led to the altered expression of the Col6a4 gene (Soret et al., 2015). Holstein mice model HSCR and present with delayed colonization of fetal bowel by ENCCs due to slower cell migration. A smaller percentage of enteric glia-fated derivatives, and a larger percentage of undifferentiated ENCCs were also observed in Holstein fetal intestine, suggesting that the mutation slows glial differentiation. The transgene insertion site maps between collagen-6α4 (Col6a4) and glycerate kinase (Glyctk). RNA-Seq analysis on isolated E12.5 ENCCs revealed that Col6a4 mRNA is markedly increased (about 250-fold) in Holstein mutants; however, collagen VI protein levels are only about 3-fold higher, an observation ascribed to the need to incorporate Col6a4 protein into trimeric collagen monomers that also contain Col6a1 and Col6a2, which are encoded by genes that are not overexpressed in mutant mice.

Other recent work suggests that post-translational regulatory mechanisms also contribute to ENS abnormalities. MicroRNAs are a class of small RNAs that bind to specific mRNA targets, leading to mRNA degradation or translational inhibition. Dicer, an RNase III endonuclease, is one of the critical enzymes for miRNA biosynthesis. Only one Dicer gene (Dicer1) exists in the mouse genome (Eppig et al., 2015), which presumably mediates the processing of all miRNAs. In order to reveal miRNA function in neural crest development, a tissue specific Dicer knockout mouse was generated through crosses with Wnt1-Cre (Huang et al., 2010; Nie et al., 2011; Zehir et al., 2010). Neural crest-restricted deletion of Dicer1 is perinatal lethal with mutants exhibiting severe defects of the craniofacial skeleton as well as the enteric, sensory and sympathetic systems. Deletion of Dicer1 does not affect neural crest cell migration and target tissue colonization; however, the post-migratory neural crest derivatives are dependent on Dicer for survival. In the ENS Dicer and the poorly characterized miRNAs it processes, are required to prevent apoptosis just before and during differentiation of ENCCs. As a result, no defects in ENS development were apparent before E14.5, but loss of Dicer dramatically decreased ENS cell density along the whole length of the fetal intestine from E17. Huang et al also demonstrated that miRNAs were required for the differentiation and survival of dopaminergic neurons in other parts of the peripheral nervous system, but detailed analysis of how Dicer loss leads to ENS deficits remains to be investigated further. To date, very few studies have investigated the roles of miRNAs in HSCR. To our knowledge, fewer than ten studies (most of them from the same group) have been published on this topic. In 2014, Tang W et al. collected close to 100 serum samples from HSCR cases and matched controls, and an initial screening of genes within which miRNAs reside or downstream regulated mRNAs was performed. Alterations of Ret, Pten, Diexf, CD47/Cul3, Sox9, and Nid1 were reported (Lei et al., 2014; Li et al., 2014; Mi et al., 2014; Sharan et al., 2015; Tang et al., 2014; Zhu et al., 2015). However, no specific deletion/mutation of miRNA leading to aganglionosis in mouse models has been reported to date. The exact role of miRNA and Dicer on major ENS players’ regulation therefore needs further investigation.

Most recently the effects of environmental factors were explicitly shown to contribute to HSCR susceptibility. To test the hypothesis that certain metabolites of common medicines might increase HSCR risk, Heuckeroth’s team treated cell cultures and animal models with several drugs commonly used during early human pregnancy. Retinoic acid, a metabolite of vitamin A, was the first tested molecule shown to influence the proliferation and differentiation of mouse ENCCs in vitro (Sato and Heuckeroth, 2008). Subsequently, retinol binding protein 4 (Rbp4−/−) mutants depleted of Vitamin A were shown to present with colorectal aganglionosis due to impaired lamellipodia formation and reduced ENCCs migration in response to Gdnf. This led to the suggestion that Vitamin A deficiency may be a non-genetic risk factor that increases HSCR penetrance and expressivity (Fu et al., 2010). Consistent with this possibility, retinaldehyde dehydrogenase (Raldh2) mutant mice typically die by E9.5, but viability can be prolonged by exogenous RA supplementation. However even after supplementation in utero with Vitamin A Raldh2−/− mutants still lack ENCCs in the fetal intestine, a finding that shows the extreme sensitivity of vagal neural crest to levels of Vitamin A (Niederreither et al., 2003). Combinatorial inactivation of the three Raldh enzymes (Raldh1−/−, Raldh2+/−, Raldh3+/−) identified reduced neuron density and altered ratios of myenteric neurons to glial in the colons of Raldh1KO, Raldh2Het, Raldh3Het mutants compared to the wild type mice. It was also found that Raldh mutants have altered colonic motility in response to mucosal stimulation. These findings indicate that each of the Raldh genes contribute to ENS development and function (Wright-Jin et al., 2013). Generation of conditional knock-outs are now needed to confirm the tissue specific requirement of retinoic acid during ENS development.

The immunosuppressant mycophenolic acid (MPA) was also shown to induce aganglionosis in mice, and enhances the penetrance and phenotype severity of mutations that model HSCR (Lake et al., 2013). In culture, MPA drastically decreases Ret+ cell migration out of bowel explants, by decreasing the percentage of migrating ENCCs with lamellipodia, reducing DNA synthesis and by inducing ENCCs apoptosis. In vivo, MPA treatment induces bowel aganglionosis and increases the penetrance and severity of aganglionosis in Ret9 and Sox10lacZ mutant mice. MPA blocks the rate-limiting step of de novo guanine nucleotide synthesis by inhibiting inosine 5’ monophosphate dehydrogenase (IMPDH), a ubiquitous metabolic enzyme whose expression is relatively enriched in ENCCs. To further explore the role for this basic metabolic pathway, Lake et al. recently deleted Impdh2 using a conditional allele crossed with the Wnt1-Cre transgene, and observed defects in multiple neural crest derivatives, including highly penetrant intestinal aganglionosis, agenesis of the craniofacial skeleton, and cardiac outflow tract with great vessel malformations (Lake et al., 2016). ENS defects in mutants lacking Impdh2 within the neural crest are clearly visible from E13.5 and demonstrate a critical role for de novo guanine nucleotide biosynthesis in ENS development. These exciting findings in mouse models suggest that some cases of HSCR may be preventable through dietary supplementation.

Most recently, Ibuprofen was shown to decrease migration and inhibit bowel colonization by ENCCs in zebrafish, chick and mouse (Schill et al., 2016). Ibuprofen treated ENCCs exhibit reduced migration, fewer lamellipodia and lower levels of active Rac1/cdc42. Additionally, inhibiting ROCK, a RhoA effector and known Rac1 antagonist, reverses Ibuprofen effects on migrating mouse ENCCs in culture. It also inhibits colonization of Ret+/− mouse bowel by ENCCs in vivo, but mice deficient in Ptgs1 (COX 1) and Ptgs2 (COX 2) show normal bowel colonization, suggesting COX-independent effects. These findings raise the concern that Ibuprofen may increase HSCR risk in some genetically susceptible families.

In addition to identifying critical molecules for ENS development, study of mouse models has also recently clarified several fundamental developmental processes that occur in ENS formation. The migratory path taken by ENCCs to colonize the gut is one example. Identification of trans-mesenteric migrating cells was revealed by engineered mouse models expressing a photo-convertible reporter that enabled live cell imaging of ENCC by live-cell video microscopy (Nishiyama et al., 2012). Such time-lapse imaging showed that ENCCs reach the hindgut by migrating though the mesentery, effectively cutting across between the midgut and the hindgut, as opposed to a migrating along the full length of the fetal intestine around the loop from midgut to caecum and then on to hindgut. The interdependent relationship between vasculature and ENCCs studies also benefited from analysis of mouse models. The vasculature and nervous system indeed share striking similarities in their networked, tree-like architecture and in the way they are super-imposed in mature organs. It has previously been suggested that the intestinal microvasculature network directs the migration of ENCCs along the gut to promote the formation of the ENS based on experiments in the chick (Nagy et al., 2009). However, fate-mapping and intravascular dye injection recently revealed that in early development both networks form independently of each other and that blood vessel networks are not necessary to guide migrating ENCCs during mouse ENS development (Delalande et al., 2014). As mentioned above, new cell populations contributing to ENS development have also been recently described (Uesaka et al., 2015). Interestingly, Schwann cell precursors (SCPs) give rise to up to 20% of enteric neurons in the large intestine under physiological conditions, and these cells give rise mainly to calretinin-expressing neurons. Another developmental aspect that was clarified is the notion of the “time window” during which the fetal bowel can be colonized. Indeed, grafting experiments previously suggested that if ENCCs have not finished migration by E14.5, aganglionosis will result. However, continued colonization has been observed between E14.5 and E18.5 in Tcof1 mutants, arguing that the delayed migration is not always predictive or sufficient for the pathogenesis of aganglionosis (Barlow et al., 2012). Rather, a balance between ENCCs proliferation in concert with differentiation and extrinsic gut micro-environmental influences are required to complete ENS formation. Another interpretation of these data could rely on the specific defects observed in Tcof1 mutants, where unlike most other genes, Tcof1 haploinsufficiency affects pre-migratory neural crest cells in the neural tube and not during migration. Hence, ENCCs seem to sense their own density and normally have considerable compensatory ability, but ENCCs with mutations in genes affecting proliferation or in maintaining progenitors in an uncommitted state, are unable to increase their rate of proliferation sufficiently to compensate when density falls below critical levels (hypothesized by Obermayr et al, 2013).

Gene Interactions that Impact the ENS

Because ENS development is highly complex, ENCCs must concurrently receive and respond to the activity of multiple factors. Initial evidence from human genetic studies suggested that interactions between variants at the RET and EDNRB genes in some HSCR patients could influence disease severity of some HSCR patients (Auricchio et al., 1999; Carrasquillo et al., 2002). The initial genetic association studies in human patients were confirmed by subsequent crosses between Ret knockout and Ednrb hypomorphic mutant mice that resulted in increased severity and penetrance of aganglionosis in Ret:Ednrb double mutants (Carrasquillo et al., 2002). Interaction between the Ret and Ednrb pathways in mice was further supported by altered ENS phenotypes in Edn3:Ret double mutants relative to single mutant animals (Barlow et al., 2003). Subsequently, combined effects between the ligands Gdnf and Edn3 in vitro were documented. While both Gdnf and Edn3 act synergistically to enhance ENCCs proliferation, they have antagonistic roles with respect to differentiation and migration with Edn3 inhibiting Gdnf-mediated effects (Barlow et al., 2003). These studies have illustrated how the coordinated activity between these two pathways is essential, particularly in the cecum where Gdnf is strongly up-regulated at the stage ENCCs begin populating this region of the intestine.

Additional interactions between signaling pathways and molecules that may underlie the variability of aganglionosis seen in HSCR patients have been identified by examining ENS deficits in double mutant offspring produced by crossing distinct single gene mouse mutants (Table 4 and references therein). Not surprisingly given the prominent expression of Sox10 early in ENCC development, multiple genes have been identified that interact with this transcription factor-encoding gene. Genes that have been found to interact with Sox10 via double mutant crosses include Ednrb, Edn3, Sox8, Zeb2, Itgb1, L1Cam, Ets1, and Sufu. Separate analysis in mouse mutants supports interactions between EdnrB and L1Cam, Itgb1 and Cdh2, as well as Pax3 and Tcof1 or Cdx. Some of these genetic interactions may be the result of direct crosstalk between the individual genes as indicated by in vitro studies showing that Sox10 binds to Ret and Ednrb regulatory regions (Lang and Epstein, 2003; Zhu et al., 2004). Other gene interactions that lead to increased aganglionosis may be indirect, and are likely mediated by common downstream signaling molecules, such as protein kinase A in the case of Ret and Ednrb interaction. These types of crosses aid in understanding the coordinated function of each molecule tested, particularly when careful detailed studies of multiple time points in ENCC migration and differentiation are analyzed.

While double mutant crosses can identify genes that exacerbate aganglionosis in HSCR models, they require knowledge of the pathway under investigation and availability of mouse mutants in specific genes. In contrast genome-wide linkage studies have the potential to identify genes that might be missed by candidate gene approaches and are capable of identifying variants that alter transcription, protein levels, or regulatory RNAs. To date there has been only a single modifier screen to identify genomic regions that influence penetrance and severity of aganglionosis in mouse models despite the ready availability of many inbred strains (Owens et al., 2005). Using Sox10Dom mutant mice bred onto distinct inbred genetic backgrounds Southard-Smith’s team identified five genomic regions that alter aganglionosis in an unbiased genome-wide survey. Because this effort did not introduce any new mutations into the genome, as is done in ENU mutagenesis, the intervals identified harbor naturally occurring variants that interact with the initial Sox10Dom allele to increase the severity or penetrance of aganglionosis in this HSCR model. Since modifier intervals are so large future fine mapping is needed to determine whether the regions on mouse chromosomes 3, 5, 8, 11 and 14 in this study correspond to previously known HSCR susceptibility genes or are indications of novel genes that remain to be identified. Developmental studies performed by the same group determined that differences in the genetic backgrounds of C57BL6J and C3HeB/FeJ strains not only altered migration of ENCCs in the fetal bowel of Sox10Dom mutants but also affected the developmental potential of ENS progenitors (Walters et al., 2010). While these studies provide insight into the genome regions that influence the Sox10Dom allele, the identified modifier genes will differ depending on the initial HSCR mutation incorporated into crosses (Sox10 versus Ednrb versus Ret). As a result such approaches offer fertile ground for identifying naturally occurring variants that influence ENS development and maturation.

A large number of candidate genes that are associated with HSCR disease susceptibility have been identified by comparative genome hybridization (Jiang et al., 2011). A number of these genes, such as the Semaphorins, have not yet been implicated as causative for aganglionosis in mouse models despite the fact that knock down experiments in zebrafish suggest an interaction between Ret and Sema3C/3D (Jiang et al., 2015). These differences may be the result of lethality in haploinsufficient Sema mouse mutants that complicates analysis of vagal ENCCs in double mutant crosses (Feiner et al., 2001). Further investigation with carefully designed mouse alleles that reduce gene expression in specific neural crest lineages will enable validation of the gene interactions detected in HSCR patients and future mechanistic analysis of the gene interactions in double mutant mouse crosses.

Conclusion

Mouse models have provided un-precedented insight into the genes and developmental processes that are essential for formation of the elaborate neural network that makes up the ENS in the bowel. Not only have single gene mutations elaborated our understanding of individual gene function, but double mutant and modifier studies of mouse HSCR models also provide a means to investigate the genetic basis of oligogenic inheritance in HSCR. New HSCR susceptibility genes are likely to emerge from ongoing mutagenesis screens and emerging genetic resources like the collaborative cross and diversity outbred lines (Churchill et al., 2004; Churchill et al., 2012; Threadgill et al., 2002). While important advances have been made in understanding the genetic basis of HSCR, it is clear that mouse models will continue to aid in identifying the genes and developmental mechanisms that lead to this complex disorder.

Highlights.

  • The enteric nervous system (ENS) develops from neural crest progenitors.

  • Mouse models have identified many genes that participate in ENS development.

  • Single gene mutations in mice can lead to aganglionosis, as in Hirschsprung disease, or cause subtle ENS alterations that affect gut motility.

  • Crosses between mouse mutants have definitively identified genetic interactions that lead to ENS deficits.

  • Development of new mouse alleles and mutant screens offer a means to understand developmental processes that impact the ENS.

Acknowledgments

The work performed by the N.B. team is supported by INSERM, Agence Nationale de la Recherche (ANR-12-BSV2-0019) and fondation ARC pour la recherche sur le cancer. Research performed by the E.M.S2. team was supported by NIH R01 DK60047 and March of Dimes FY12-450.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Disclosures

Authors declare no conflict of interest.

Contributor Information

Nadege Bondurand, Email: nadege.bondurand@inserm.fr.

E. Michelle Southard-Smith, Email: Michelle.southard-smith@vanderbilt.edu.

References

  1. Akbareian SE, Nagy N, Steiger CE, Mably JD, Miller SA, Hotta R, Molnar D, Goldstein AM. Enteric neural crest-derived cells promote their migration by modifying their microenvironment through tenascin-C production. Dev Biol. 2013;382:446–456. doi: 10.1016/j.ydbio.2013.08.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Amiel J, Laudier B, Attie-Bitach T, Trang H, de Pontual L, Gener B, Trochet D, Etchevers H, Ray P, Simonneau M, Vekemans M, Munnich A, Gaultier C, Lyonnet S. Polyalanine expansion and frameshift mutations of the paired-like homeobox gene PHOX2B in congenital central hypoventilation syndrome. Nat Genet. 2003;33:459–461. doi: 10.1038/ng1130. [DOI] [PubMed] [Google Scholar]
  3. Amiel J, Sproat-Emison E, Garcia-Barcelo M, Lantieri F, Burzynski G, Borrego S, Pelet A, Arnold S, Miao X, Griseri P, Brooks AS, Antinolo G, de Pontual L, Clement-Ziza M, Munnich A, Kashuk C, West K, Wong KK, Lyonnet S, Chakravarti A, Tam PK, Ceccherini I, Hofstra RM, Fernandez R, Hirschsprung Disease C. Hirschsprung disease, associated syndromes and genetics: a review. J Med Genet. 2008;45:1–14. doi: 10.1136/jmg.2007.053959. [DOI] [PubMed] [Google Scholar]
  4. Anderson RB, Turner KN, Nikonenko AG, Hemperly J, Schachner M, Young HM. The cell adhesion molecule l1 is required for chain migration of neural crest cells in the developing mouse gut. Gastroenterology. 2006;130:1221–1232. doi: 10.1053/j.gastro.2006.01.002. [DOI] [PubMed] [Google Scholar]
  5. Angrist M, Bolk S, Halushka M, Lapchak PA, Chakravarti A. Germline mutations in glial cell line-derived neurotrophic factor (GDNF) and RET in a Hirschsprung disease patient. Nat Genet. 1996;14:341–344. doi: 10.1038/ng1196-341. [DOI] [PubMed] [Google Scholar]
  6. Antonellis A, Bennett WR, Menheniott TR, Prasad AB, Lee-Lin SQ, Program NCS, Green ED, Paisley D, Kelsh RN, Pavan WJ, Ward A. Deletion of long-range sequences at Sox10 compromises developmental expression in a mouse model of Waardenburg-Shah (WS4) syndrome. Hum Mol Genet. 2006;15:259–271. doi: 10.1093/hmg/ddi442. [DOI] [PubMed] [Google Scholar]
  7. Asai N, Fukuda T, Wu Z, Enomoto A, Pachnis V, Takahashi M, Costantini F. Targeted mutation of serine 697 in the Ret tyrosine kinase causes migration defect of enteric neural crest cells. Development. 2006;133:4507–4516. doi: 10.1242/dev.02616. [DOI] [PubMed] [Google Scholar]
  8. Attie T, Pelet A, Edery P, Eng C, Mulligan LM, Amiel J, Boutrand L, Beldjord C, Nihoul-Fekete C, Munnich A, et al. Diversity of RET proto-oncogene mutations in familial and sporadic Hirschsprung disease. Hum Mol Genet. 1995;4:1381–1386. doi: 10.1093/hmg/4.8.1381. [DOI] [PubMed] [Google Scholar]
  9. Attie T, Pelet A, Sarda P, Eng C, Edery P, Mulligan LM, Ponder BA, Munnich A, Lyonnet S. A 7 bp deletion of the RET proto-oncogene in familial Hirschsprung’s disease. Hum Mol Genet. 1994;3:1439–1440. doi: 10.1093/hmg/3.8.1439. [DOI] [PubMed] [Google Scholar]
  10. Auricchio A, Griseri P, Carpentieri ML, Betsos N, Staiano A, Tozzi A, Priolo M, Thompson H, Bocciardi R, Romeo G, Ballabio A, Ceccherini I. Double heterozygosity for a RET substitution interfering with splicing and an EDNRB missense mutation in Hirschsprung disease. Am J Hum Genet. 1999;64:1216–1221. doi: 10.1086/302329. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Badner JA, Sieber WK, Garver KL, Chakravarti A. A genetic study of Hirschsprung disease. Am J Hum Genet. 1990;46:568–580. [PMC free article] [PubMed] [Google Scholar]
  12. Barlow A, de Graaff E, Pachnis V. Enteric nervous system progenitors are coordinately controlled by the G protein-coupled receptor EDNRB and the receptor tyrosine kinase RET. Neuron. 2003;40:905–916. doi: 10.1016/s0896-6273(03)00730-x. [DOI] [PubMed] [Google Scholar]
  13. Barlow AJ, Dixon J, Dixon M, Trainor PA. Tcof1 acts as a modifier of Pax3 during enteric nervous system development and in the pathogenesis of colonic aganglionosis. Hum Mol Genet. 2013;22:1206–1217. doi: 10.1093/hmg/dds528. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Barlow AJ, Dixon J, Dixon MJ, Trainor PA. Balancing neural crest cell intrinsic processes with those of the microenvironment in Tcof1 haploinsufficient mice enables complete enteric nervous system formation. Hum Mol Genet. 2012;21:1782–1793. doi: 10.1093/hmg/ddr611. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Bates MD, Dunagan DT, Welch LC, Kaul A, Harvey RP. The Hlx homeobox transcription factor is required early in enteric nervous system development. BMC Dev Biol. 2006;6:33. doi: 10.1186/1471-213X-6-33. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Baynash AG, Hosoda K, Giaid A, Richardson JA, Emoto N, Hammer RE, Yanagisawa M. Interaction of endothelin-3 with endothelin-B receptor is essential for development of epidermal melanocytes and enteric neurons. Cell. 1994;79:1277–1285. doi: 10.1016/0092-8674(94)90018-3. [DOI] [PubMed] [Google Scholar]
  17. Bergeron KF, Cardinal T, Toure AM, Beland M, Raiwet DL, Silversides DW, Pilon N. Male-biased aganglionic megacolon in the TashT mouse line due to perturbation of silencer elements in a large gene desert of chromosome 10. PLoS Genet. 2015;11:e1005093. doi: 10.1371/journal.pgen.1005093. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Bergeron KF, Silversides DW, Pilon N. The developmental genetics of Hirschsprung’s disease. Clin Genet. 2013;83:15–22. doi: 10.1111/cge.12032. [DOI] [PubMed] [Google Scholar]
  19. Biau S, Jin S, Fan CM. Gastrointestinal defects of the Gas1 mutant involve dysregulated Hedgehog and Ret signaling. Biol Open. 2013;2:144–155. doi: 10.1242/bio.20123186. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Blaugrund E, Pham TD, Tennyson VM, Lo L, Sommer L, Anderson DJ, Gershon MD. Distinct subpopulations of enteric neuronal progenitors defined by time of development, sympathoadrenal lineage markers and Mash-1-dependence. Development. 1996;122:309–320. doi: 10.1242/dev.122.1.309. [DOI] [PubMed] [Google Scholar]
  21. Bondurand N, Natarajan D, Barlow A, Thapar N, Pachnis V. Maintenance of mammalian enteric nervous system progenitors by SOX10 and endothelin 3 signalling. Development. 2006;133:2075–2086. doi: 10.1242/dev.02375. [DOI] [PubMed] [Google Scholar]
  22. Bondurand N, Sham MH. The role of SOX10 during enteric nervous system development. Dev Biol. 2013;382:330–343. doi: 10.1016/j.ydbio.2013.04.024. [DOI] [PubMed] [Google Scholar]
  23. Breau MA, Dahmani A, Broders-Bondon F, Thiery JP, Dufour S. Beta1 integrins are required for the invasion of the caecum and proximal hindgut by enteric neural crest cells. Development. 2009;136:2791–2801. doi: 10.1242/dev.031419. [DOI] [PubMed] [Google Scholar]
  24. Breau MA, Pietri T, Eder O, Blanche M, Brakebusch C, Fassler R, Thiery JP, Dufour S. Lack of beta1 integrins in enteric neural crest cells leads to a Hirschsprung-like phenotype. Development. 2006;133:1725–1734. doi: 10.1242/dev.02346. [DOI] [PubMed] [Google Scholar]
  25. Britsch S, Goerich DE, Riethmacher D, Peirano RI, Rossner M, Nave KA, Birchmeier C, Wegner M. The transcription factor Sox10 is a key regulator of peripheral glial development. Genes Dev. 2001;15:66–78. doi: 10.1101/gad.186601. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Broders-Bondon F, Paul-Gilloteaux P, Carlier C, Radice GL, Dufour S. N-cadherin and beta1-integrins cooperate during the development of the enteric nervous system. Dev Biol. 2012;364:178–191. doi: 10.1016/j.ydbio.2012.02.001. [DOI] [PubMed] [Google Scholar]
  27. Brooks AS, Oostra BA, Hofstra RM. Studying the genetics of Hirschsprung’s disease: unraveling an oligogenic disorder. Clin Genet. 2005;67:6–14. doi: 10.1111/j.1399-0004.2004.00319.x. [DOI] [PubMed] [Google Scholar]
  28. Brun P, Giron MC, Qesari M, Porzionato A, Caputi V, Zoppellaro C, Banzato S, Grillo AR, Spagnol L, De Caro R, Pizzuti D, Barbieri V, Rosato A, Sturniolo GC, Martines D, Zaninotto G, Palu G, Castagliuolo I. Toll-like receptor 2 regulates intestinal inflammation by controlling integrity of the enteric nervous system. Gastroenterology. 2013;145:1323–1333. doi: 10.1053/j.gastro.2013.08.047. [DOI] [PubMed] [Google Scholar]
  29. Burns AJ, Douarin NM. The sacral neural crest contributes neurons and glia to the post-umbilical gut: spatiotemporal analysis of the development of the enteric nervous system. Development. 1998;125:4335–4347. doi: 10.1242/dev.125.21.4335. [DOI] [PubMed] [Google Scholar]
  30. Butler Tjaden NE, Trainor PA. The developmental etiology and pathogenesis of Hirschsprung disease. Transl Res. 2013;162:1–15. doi: 10.1016/j.trsl.2013.03.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Cacalano G, Farinas I, Wang LC, Hagler K, Forgie A, Moore M, Armanini M, Phillips H, Ryan AM, Reichardt LF, Hynes M, Davies A, Rosenthal A. GFRalpha1 is an essential receptor component for GDNF in the developing nervous system and kidney. Neuron. 1998;21:53–62. doi: 10.1016/s0896-6273(00)80514-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Cantrell VA, Owens SE, Chandler RL, Airey DC, Bradley KM, Smith JR, Southard-Smith EM. Interactions between Sox10 and EdnrB modulate penetrance and severity of aganglionosis in the Sox10Dom mouse model of Hirschsprung disease. Hum Mol Genet. 2004;13:2289–2301. doi: 10.1093/hmg/ddh243. [DOI] [PubMed] [Google Scholar]
  33. Carniti C, Belluco S, Riccardi E, Cranston AN, Mondellini P, Ponder BA, Scanziani E, Pierotti MA, Bongarzone I. The Ret(C620R) mutation affects renal and enteric development in a mouse model of Hirschsprung’s disease. Am J Pathol. 2006;168:1262–1275. doi: 10.2353/ajpath.2006.050607. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Carrasquillo MM, McCallion AS, Puffenberger EG, Kashuk CS, Nouri N, Chakravarti A. Genome-wide association study and mouse model identify interaction between RET and EDNRB pathways in Hirschsprung disease. Nat Genet. 2002;32:237–244. doi: 10.1038/ng998. [DOI] [PubMed] [Google Scholar]
  35. Chakravarti A, McCallion AS, Lyonnet S. Chapter 251: Hirschsprung Disease., The Metabolic and Molecular Bases of Inherited Disease. The McGraw-Hill Companies. 2004:1–47. [Google Scholar]
  36. Chalazonitis A, D’Autreaux F, Guha U, Pham TD, Faure C, Chen JJ, Roman D, Kan L, Rothman TP, Kessler JA, Gershon MD. Bone morphogenetic protein-2 and −4 limit the number of enteric neurons but promote development of a TrkC-expressing neurotrophin-3-dependent subset. J Neurosci. 2004;24:4266–4282. doi: 10.1523/JNEUROSCI.3688-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Chalazonitis A, D’Autreaux F, Pham TD, Kessler JA, Gershon MD. Bone morphogenetic proteins regulate enteric gliogenesis by modulating ErbB3 signaling. Dev Biol. 2011a;350:64–79. doi: 10.1016/j.ydbio.2010.11.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Chalazonitis A, Pham TD, Li Z, Roman D, Guha U, Gomes W, Kan L, Kessler JA, Gershon MD. Bone morphogenetic protein regulation of enteric neuronal phenotypic diversity: relationship to timing of cell cycle exit. J Comp Neurol. 2008;509:474–492. doi: 10.1002/cne.21770. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Chalazonitis A, Pham TD, Rothman TP, DiStefano PS, Bothwell M, Blair-Flynn J, Tessarollo L, Gershon MD. Neurotrophin-3 is required for the survival-differentiation of subsets of developing enteric neurons. J Neurosci. 2001;21:5620–5636. doi: 10.1523/JNEUROSCI.21-15-05620.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Chalazonitis A, Tang AA, Shang Y, Pham TD, Hsieh I, Setlik W, Gershon MD, Huang EJ. Homeodomain interacting protein kinase 2 regulates postnatal development of enteric dopaminergic neurons and glia via BMP signaling. J Neurosci. 2011b;31:13746–13757. doi: 10.1523/JNEUROSCI.1078-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Churchill GA, Airey DC, Allayee H, Angel JM, Attie AD, Beatty J, Beavis WD, Belknap JK, Bennett B, Berrettini W, Bleich A, Bogue M, Broman KW, Buck KJ, Buckler E, Burmeister M, Chesler EJ, Cheverud JM, Clapcote S, Cook MN, Cox RD, Crabbe JC, Crusio WE, Darvasi A, Deschepper CF, Doerge RW, Farber CR, Forejt J, Gaile D, Garlow SJ, Geiger H, Gershenfeld H, Gordon T, Gu J, Gu W, de Haan G, Hayes NL, Heller C, Himmelbauer H, Hitzemann R, Hunter K, Hsu HC, Iraqi FA, Ivandic B, Jacob HJ, Jansen RC, Jepsen KJ, Johnson DK, Johnson TE, Kempermann G, Kendziorski C, Kotb M, Kooy RF, Llamas B, Lammert F, Lassalle JM, Lowenstein PR, Lu L, Lusis A, Manly KF, Marcucio R, Matthews D, Medrano JF, Miller DR, Mittleman G, Mock BA, Mogil JS, Montagutelli X, Morahan G, Morris DG, Mott R, Nadeau JH, Nagase H, Nowakowski RS, O’Hara BF, Osadchuk AV, Page GP, Paigen B, Paigen K, Palmer AA, Pan HJ, Peltonen-Palotie L, Peirce J, Pomp D, Pravenec M, Prows DR, Qi Z, Reeves RH, Roder J, Rosen GD, Schadt EE, Schalkwyk LC, Seltzer Z, Shimomura K, Shou S, Sillanpaa MJ, Siracusa LD, Snoeck HW, Spearow JL, Svenson K, Tarantino LM, Threadgill D, Toth LA, Valdar W, de Villena FP, Warden C, Whatley S, Williams RW, Wiltshire T, Yi N, Zhang D, Zhang M, Zou F, Complex Trait C. The Collaborative Cross, a community resource for the genetic analysis of complex traits. Nat Genet. 2004;36:1133–1137. doi: 10.1038/ng1104-1133. [DOI] [PubMed] [Google Scholar]
  42. Churchill GA, Gatti DM, Munger SC, Svenson KL. The Diversity Outbred mouse population. Mamm Genome. 2012;23:713–718. doi: 10.1007/s00335-012-9414-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Corpening JC, Deal KK, Cantrell VA, Skelton SB, Buehler DP, Southard-Smith EM. Isolation and live imaging of enteric progenitors based on Sox10-Histone2BVenus transgene expression. Genesis. 2011;49:599–618. doi: 10.1002/dvg.20748. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Coventry S, Yost C, Palmiter RD, Kapur RP. Migration of ganglion cell precursors in the ileoceca of normal and lethal spotted embryos, a murine model for Hirschsprung disease. Lab Invest. 1994;71:82–93. [PubMed] [Google Scholar]
  45. Crone SA, Negro A, Trumpp A, Giovannini M, Lee KF. Colonic epithelial expression of ErbB2 is required for postnatal maintenance of the enteric nervous system. Neuron. 2003;37:29–40. doi: 10.1016/s0896-6273(02)01128-5. [DOI] [PubMed] [Google Scholar]
  46. D’Autreaux F, Margolis KG, Roberts J, Stevanovic K, Mawe G, Li Z, Karamooz N, Ahuja A, Morikawa Y, Cserjesi P, Setlick W, Gershon MD. Expression level of Hand2 affects specification of enteric neurons and gastrointestinal function in mice. Gastroenterology. 2011;141:576–587. 587, e571–e576. doi: 10.1053/j.gastro.2011.04.059. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. de Graaff E, Srinivas S, Kilkenny C, D’Agati V, Mankoo BS, Costantini F, Pachnis V. Differential activities of the RET tyrosine kinase receptor isoforms during mammalian embryogenesis. Genes Dev. 2001;15:2433–2444. doi: 10.1101/gad.205001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Delalande JM, Natarajan D, Vernay B, Finlay M, Ruhrberg C, Thapar N, Burns AJ. Vascularisation is not necessary for gut colonisation by enteric neural crest cells. Dev Biol. 2014;385:220–229. doi: 10.1016/j.ydbio.2013.11.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Druckenbrod NR, Epstein ML. The pattern of neural crest advance in the tenascin-C and colon. Dev Biol. 2005;287:125–133. doi: 10.1016/j.ydbio.2005.08.040. [DOI] [PubMed] [Google Scholar]
  50. Druckenbrod NR, Epstein ML. Age-dependent changes in the gut environment restrict the invasion of the hindgut by enteric neural progenitors. Development. 2009;136:3195–3203. doi: 10.1242/dev.031302. [DOI] [PubMed] [Google Scholar]
  51. Durbec P, Marcos-Gutierrez CV, Kilkenny C, Grigoriou M, Wartiowaara K, Suvanto P, Smith D, Ponder B, Costantini F, Saarma M, et al. GDNF signalling through the Ret receptor tyrosine kinase. Nature. 1996;381:789–793. doi: 10.1038/381789a0. [DOI] [PubMed] [Google Scholar]
  52. Edery P, Lyonnet S, Mulligan LM, Pelet A, Dow E, Abel L, Holder S, Nihoul-Fekete C, Ponder BA, Munnich A. Mutations of the RET proto-oncogene in Hirschsprung’s disease. Nature. 1994;367:378–380. doi: 10.1038/367378a0. [DOI] [PubMed] [Google Scholar]
  53. Eketjall S, Ibanez CF. Functional characterization of mutations in the GDNF gene of patients with Hirschsprung disease. Hum Mol Genet. 2002;11:325–329. doi: 10.1093/hmg/11.3.325. [DOI] [PubMed] [Google Scholar]
  54. Emison ES, McCallion AS, Kashuk CS, Bush RT, Grice E, Lin S, Portnoy ME, Cutler DJ, Green ED, Chakravarti A. A common sex-dependent mutation in a RET enhancer underlies Hirschsprung disease risk. Nature. 2005;434:857–863. doi: 10.1038/nature03467. [DOI] [PubMed] [Google Scholar]
  55. Enomoto H, Araki T, Jackman A, Heuckeroth RO, Snider WD, Johnson EM, Jr, Milbrandt J. GFR alpha1-deficient mice have deficits in the enteric nervous system and kidneys. Neuron. 1998;21:317–324. doi: 10.1016/s0896-6273(00)80541-3. [DOI] [PubMed] [Google Scholar]
  56. Eppig JT, Blake JA, Bult CJ, Kadin JA, Richardson JE, Mouse Genome Database G. The Mouse Genome Database (MGD): facilitating mouse as a model for human biology and disease. Nucleic Acids Res. 2015;43:D726–D736. doi: 10.1093/nar/gku967. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Feiner L, Webber AL, Brown CB, Lu MM, Jia L, Feinstein P, Mombaerts P, Epstein JA, Raper JA. Targeted disruption of semaphorin 3C leads to persistent truncus arteriosus and aortic arch interruption. Development. 2001;128:3061–3070. doi: 10.1242/dev.128.16.3061. [DOI] [PubMed] [Google Scholar]
  58. Fu M, Landreville S, Agapova OA, Wiley LA, Shoykhet M, Harbour JW, Heuckeroth RO. Retinoblastoma protein prevents enteric nervous system defects and intestinal pseudo-obstruction. J Clin Invest. 2013;123:5152–5164. doi: 10.1172/JCI67653. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Fu M, Sato Y, Lyons-Warren A, Zhang B, Kane MA, Napoli JL, Heuckeroth RO. Vitamin A facilitates enteric nervous system precursor migration by reducing Pten accumulation. Development. 2010;137:631–640. doi: 10.1242/dev.040550. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Fuchs S, Herzog D, Sumara G, Buchmann-Moller S, Civenni G, Wu X, Chrostek-Grashoff A, Suter U, Ricci R, Relvas JB, Brakebusch C, Sommer L. Stage-specific control of neural crest stem cell proliferation by the small rho GTPases Cdc42 and Rac1. Cell Stem Cell. 2009;4:236–247. doi: 10.1016/j.stem.2009.01.017. [DOI] [PubMed] [Google Scholar]
  61. Fujimoto T. Natural history and pathophysiology of enterocolitis in the piebald lethal mouse model of Hirschsprung’s disease. J Pediatr Surg. 1988;23:237–242. doi: 10.1016/s0022-3468(88)80730-9. [DOI] [PubMed] [Google Scholar]
  62. Furness JB. The enteric nervous system and neurogastroenterology. Nat Rev Gastroenterol Hepatol. 2012;9:286–294. doi: 10.1038/nrgastro.2012.32. [DOI] [PubMed] [Google Scholar]
  63. Gianino S, Grider JR, Cresswell J, Enomoto H, Heuckeroth RO. GDNF availability determines enteric neuron number by controlling precursor proliferation. Development. 2003;130:2187–2198. doi: 10.1242/dev.00433. [DOI] [PubMed] [Google Scholar]
  64. Goldstein AM, Hofstra RM, Burns AJ. Building a brain in the gut: development of the enteric nervous system. Clin Genet. 2013;83:307–316. doi: 10.1111/cge.12054. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Griseri P, Lantieri F, Puppo F, Bachetti T, Di Duca M, Ravazzolo R, Ceccherini I. A common variant located in the 3’UTR of the RET gene is associated with protection from Hirschsprung disease. Hum Mutat. 2007;28:168–176. doi: 10.1002/humu.20397. [DOI] [PubMed] [Google Scholar]
  66. Hagl CI, Klotz M, Wink E, Kranzle K, Holland-Cunz S, Gretz N, Diener M, Schafer KH. Temporal and regional morphological differences as a consequence of FGF-2 deficiency are mirrored in the myenteric proteome. Pediatr Surg Int. 2008;24:49–60. doi: 10.1007/s00383-007-2041-4. [DOI] [PubMed] [Google Scholar]
  67. Hagl CI, Wink E, Scherf S, Heumuller-Klug S, Hausott B, Schafer KH. FGF2 deficit during development leads to specific neuronal cell loss in the enteric nervous system. Histochem Cell Biol. 2013;139:47–57. doi: 10.1007/s00418-012-1023-3. [DOI] [PubMed] [Google Scholar]
  68. Hansen C, Li JY. Beyond alpha-synuclein transfer: pathology propagation in Parkinson’s disease. Trends Mol Med. 2012;18:248–255. doi: 10.1016/j.molmed.2012.03.002. [DOI] [PubMed] [Google Scholar]
  69. Harrison C, Shepherd IT. Choices choices: regulation of precursor differentiation during enteric nervous system development. Neurogastroenterol Motil. 2013;25:554–562. doi: 10.1111/nmo.12142. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Hatano M, Aoki T, Dezawa M, Yusa S, Iitsuka Y, Koseki H, Taniguchi M, Tokuhisa T. A novel pathogenesis of megacolon in Ncx/Hox11L.1 deficient mice. J Clin Invest. 1997;100:795–801. doi: 10.1172/JCI119593. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Hearn CJ, Murphy M, Newgreen D. GDNF and ET-3 differentially modulate the numbers of avian enteric neural crest cells and enteric neurons in vitro. Dev Biol. 1998;197:93–105. doi: 10.1006/dbio.1998.8876. [DOI] [PubMed] [Google Scholar]
  72. Hendershot TJ, Liu H, Sarkar AA, Giovannucci DR, Clouthier DE, Abe M, Howard MJ. Expression of Hand2 is sufficient for neurogenesis and cell type-specific gene expression in the enteric nervous system. Dev Dyn. 2007;236:93–105. doi: 10.1002/dvdy.20989. [DOI] [PubMed] [Google Scholar]
  73. Herbarth B, Pingault V, Bondurand N, Kuhlbrodt K, Hermans-Borgmeyer I, Puliti A, Lemort N, Goossens M, Wegner M. Mutation of the Sry-related Sox10 gene in Dominant megacolon, a mouse model for human Hirschsprung disease. Proc Natl Acad Sci U S A. 1998;95:5161–5165. doi: 10.1073/pnas.95.9.5161. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Heuckeroth RO, Enomoto H, Grider JR, Golden JP, Hanke JA, Jackman A, Molliver DC, Bardgett ME, Snider WD, Johnson EM, Jr, Milbrandt J. Gene targeting reveals a critical role for neurturin in the development and maintenance of enteric, sensory, and parasympathetic neurons. Neuron. 1999;22:253–263. doi: 10.1016/s0896-6273(00)81087-9. [DOI] [PubMed] [Google Scholar]
  75. Heuckeroth RO, Schafer KH. Gene-environment interactions and the enteric nervous system: Neural plasticity and Hirschsprung disease prevention. Dev Biol. 2016 doi: 10.1016/j.ydbio.2016.03.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Hofstra RM, Osinga J, Tan-Sindhunata G, Wu Y, Kamsteeg EJ, Stulp RP, van Ravenswaaij-Arts C, Majoor-Krakauer D, Angrist M, Chakravarti A, Meijers C, Buys CH. A homozygous mutation in the endothelin-3 gene associated with a combined Waardenburg type 2 and Hirschsprung phenotype (Shah-Waardenburg syndrome) Nat Genet. 1996;12:445–447. doi: 10.1038/ng0496-445. [DOI] [PubMed] [Google Scholar]
  77. Hofstra RM, Valdenaire O, Arch E, Osinga J, Kroes H, Loffler BM, Hamosh A, Meijers C, Buys CH. A loss-of-function mutation in the endothelin-converting enzyme 1 (ECE-1) associated with Hirschsprung disease, cardiac defects, and autonomic dysfunction. Am J Hum Genet. 1999;64:304–308. doi: 10.1086/302184. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Hosoda K, Hammer RE, Richardson JA, Baynash AG, Cheung JC, Giaid A, Yanagisawa M. Targeted and natural (piebald-lethal) mutations of endothelin-B receptor gene produce megacolon associated with spotted coat color in mice. Cell. 1994;79:1267–1276. doi: 10.1016/0092-8674(94)90017-5. [DOI] [PubMed] [Google Scholar]
  79. Huang T, Liu Y, Huang M, Zhao X, Cheng L. Wnt1-cre-mediated conditional loss of Dicer results in malformation of the midbrain and cerebellum and failure of neural crest and dopaminergic differentiation in mice. J Mol Cell Biol. 2010;2:152–163. doi: 10.1093/jmcb/mjq008. [DOI] [PubMed] [Google Scholar]
  80. Ivanchuk SM, Myers SM, Eng C, Mulligan LM. De novo mutation of GDNF, ligand for the RET/GDNFR-alpha receptor complex, in Hirschsprung disease. Hum Mol Genet. 1996;5:2023–2026. doi: 10.1093/hmg/5.12.2023. [DOI] [PubMed] [Google Scholar]
  81. Jain S, Naughton CK, Yang M, Strickland A, Vij K, Encinas M, Golden J, Gupta A, Heuckeroth R, Johnson EM, Jr, Milbrandt J. Mice expressing a dominant-negative Ret mutation phenocopy human Hirschsprung disease and delineate a direct role of Ret in spermatogenesis. Development. 2004;131:5503–5513. doi: 10.1242/dev.01421. [DOI] [PubMed] [Google Scholar]
  82. Jannot AS, Pelet A, Henrion-Caude A, Chaoui A, Masse-Morel M, Arnold S, Sanlaville D, Ceccherini I, Borrego S, Hofstra RM, Munnich A, Bondurand N, Chakravarti A, Clerget-Darpoux F, Amiel J, Lyonnet S. Chromosome 21 scan in Down syndrome reveals DSCAM as a predisposing locus in Hirschsprung disease. PLoS One. 2013;8:e62519. doi: 10.1371/journal.pone.0062519. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Jiang Q, Arnold S, Heanue T, Kilambi KP, Doan B, Kapoor A, Ling AY, Sosa MX, Guy M, Jiang Q, Burzynski G, West K, Bessling S, Griseri P, Amiel J, Fernandez RM, Verheij JB, Hofstra RM, Borrego S, Lyonnet S, Ceccherini I, Gray JJ, Pachnis V, McCallion AS, Chakravarti A. Functional loss of semaphorin 3C and/or semaphorin 3D and their epistatic interaction with ret are critical to Hirschsprung disease liability. Am J Hum Genet. 2015;96:581–596. doi: 10.1016/j.ajhg.2015.02.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Jiang Q, Ho YY, Hao L, Nichols Berrios C, Chakravarti A. Copy number variants in candidate genes are genetic modifiers of Hirschsprung disease. PLoS One. 2011;6:e21219. doi: 10.1371/journal.pone.0021219. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Jin S, Martinelli DC, Zheng X, Tessier-Lavigne M, Fan CM. Gas1 is a receptor for sonic hedgehog to repel enteric axons. Proc Natl Acad Sci U S A. 2015;112:E73–E80. doi: 10.1073/pnas.1418629112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Kapur RP. Early death of neural crest cells is responsible for total enteric aganglionosis in Sox10(Dom)/Sox10(Dom) mouse embryos. Pediatr Dev Pathol. 1999;2:559–569. doi: 10.1007/s100249900162. [DOI] [PubMed] [Google Scholar]
  87. Kapur RP. Colonization of the murine hindgut by sacral crest-derived neural precursors: experimental support for an evolutionarily conserved model. Dev Biol. 2000;227:146–155. doi: 10.1006/dbio.2000.9886. [DOI] [PubMed] [Google Scholar]
  88. Kim J, Lo L, Dormand E, Anderson DJ. SOX10 maintains multipotency and inhibits neuronal differentiation of neural crest stem cells. Neuron. 2003;38:17–31. doi: 10.1016/s0896-6273(03)00163-6. [DOI] [PubMed] [Google Scholar]
  89. Lake JI, Avetisyan M, Zimmermann AG, Heuckeroth RO. Neural crest requires Impdh2 for development of the enteric nervous system, great vessels, and craniofacial skeleton. Dev Biol. 2016;409:152–165. doi: 10.1016/j.ydbio.2015.11.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Lake JI, Heuckeroth RO. Enteric nervous system development: migration, differentiation, and disease. Am J Physiol Gastrointest Liver Physiol. 2013;305:G1–G24. doi: 10.1152/ajpgi.00452.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Lake JI, Tusheva OA, Graham BL, Heuckeroth RO. Hirschsprung-like disease is exacerbated by reduced de novo GMP synthesis. J Clin Invest. 2013;123:4875–4887. doi: 10.1172/JCI69781. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Lang D, Chen F, Milewski R, Li J, Lu MM, Epstein JA. Pax3 is required for enteric ganglia formation and functions with Sox10 to modulate expression of c-ret. J Clin Invest. 2000;106:963–971. doi: 10.1172/JCI10828. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Lang D, Epstein JA. Sox10 and Pax3 physically interact to mediate activation of a conserved c-RET enhancer. Hum Mol Genet. 2003;12:937–945. doi: 10.1093/hmg/ddg107. [DOI] [PubMed] [Google Scholar]
  94. Le Douarin NM, Teillet MA. Experimental analysis of the migration and differentiation of neuroblasts of the autonomic nervous system and of neurectodermal mesenchymal derivatives, using a biological cell marking technique. Dev Biol. 1974;41:162–184. doi: 10.1016/0012-1606(74)90291-7. [DOI] [PubMed] [Google Scholar]
  95. Lecerf L, Kavo A, Ruiz-Ferrer M, Baral V, Watanabe Y, Chaoui A, Pingault V, Borrego S, Bondurand N. An impairment of long distance SOX10 regulatory elements underlies isolated Hirschsprung disease. Hum Mutat. 2014;35:303–307. doi: 10.1002/humu.22499. [DOI] [PubMed] [Google Scholar]
  96. Lei H, Tang J, Li H, Zhang H, Lu C, Chen H, Li W, Xia Y, Tang W. MiR-195 affects cell migration and cell proliferation by down-regulating DIEXF in Hirschsprung’s disease. BMC Gastroenterol. 2014;14:123. doi: 10.1186/1471-230X-14-123. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Lei J, Howard MJ. Targeted deletion of Hand2 in enteric neural precursor cells affects its functions in neurogenesis, neurotransmitter specification and gangliogenesis, causing functional aganglionosis. Development. 2011;138:4789–4800. doi: 10.1242/dev.060053. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Li H, Tang J, Lei H, Cai P, Zhu H, Li B, Xu X, Xia Y, Tang W. Decreased MiR-200a/141 suppress cell migration and proliferation by targeting PTEN in Hirschsprung’s disease. Cell Physiol Biochem. 2014;34:543–553. doi: 10.1159/000363021. [DOI] [PubMed] [Google Scholar]
  99. Li Z, Caron MG, Blakely RD, Margolis KG, Gershon MD. Dependence of serotonergic and other nonadrenergic enteric neurons on norepinephrine transporter expression. J Neurosci. 2010;30:16730–16740. doi: 10.1523/JNEUROSCI.2276-10.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Li Z, Chalazonitis A, Huang YY, Mann JJ, Margolis KG, Yang QM, Kim DO, Cote F, Mallet J, Gershon MD. Essential roles of enteric neuronal serotonin in gastrointestinal motility and the development/survival of enteric dopaminergic neurons. J Neurosci. 2011;31:8998–9009. doi: 10.1523/JNEUROSCI.6684-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Liu JA, Lai FP, Gui HS, Sham MH, Tam PK, Garcia-Barcelo MM, Hui CC, Ngan ES. Identification of GLI Mutations in Patients With Hirschsprung Disease That Disrupt Enteric Nervous System Development in Mice. Gastroenterology. 2015;149:1837–1848. e1835. doi: 10.1053/j.gastro.2015.07.060. [DOI] [PubMed] [Google Scholar]
  102. Lui VC, Cheng WW, Leon TY, Lau DK, Garcia-Barcelo MM, Miao XP, Kam MK, So MT, Chen Y, Wall NA, Sham MH, Tam PK. Perturbation of hoxb5 signaling in vagal neural crests down-regulates ret leading to intestinal hypoganglionosis in mice. Gastroenterology. 2008;134:1104–1115. doi: 10.1053/j.gastro.2008.01.028. [DOI] [PubMed] [Google Scholar]
  103. Maka M, Stolt CC, Wegner M. Identification of Sox8 as a modifier gene in a mouse model of Hirschsprung disease reveals underlying molecular defect. Dev Biol. 2005;277:155–169. doi: 10.1016/j.ydbio.2004.09.014. [DOI] [PubMed] [Google Scholar]
  104. McCallion AS, Stames E, Conlon RA, Chakravarti A. Phenotype variation in two-locus mouse models of Hirschsprung disease: tissue-specific interaction between Ret and Ednrb. Proc Natl Acad Sci U S A. 2003;100:1826–1831. doi: 10.1073/pnas.0337540100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. McKeown SJ, Stamp L, Hao MM, Young HM. Hirschsprung disease: a developmental disorder of the enteric nervous system. Wiley Interdiscip Rev Dev Biol. 2013;2:113–129. doi: 10.1002/wdev.57. [DOI] [PubMed] [Google Scholar]
  106. Mi J, Chen D, Wu M, Wang W, Gao H. Study of the effect of miR124 and the SOX9 target gene in Hirschsprung’s disease. Mol Med Rep. 2014;9:1839–1843. doi: 10.3892/mmr.2014.2022. [DOI] [PubMed] [Google Scholar]
  107. Moore MW, Klein RD, Farinas I, Sauer H, Armanini M, Phillips H, Reichardt LF, Ryan AM, Carver-Moore K, Rosenthal A. Renal and neuronal abnormalities in mice lacking GDNF. Nature. 1996;382:76–79. doi: 10.1038/382076a0. [DOI] [PubMed] [Google Scholar]
  108. Morikawa Y, D’Autreaux F, Gershon MD, Cserjesi P. Hand2 determines the noradrenergic phenotype in the mouse sympathetic nervous system. Dev Biol. 2007;307:114–126. doi: 10.1016/j.ydbio.2007.04.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Mundell NA, Plank JL, LeGrone AW, Frist AY, Zhu L, Shin MK, Southard-Smith EM, Labosky PA. Enteric nervous system specific deletion of Foxd3 disrupts glial cell differentiation and activates compensatory enteric progenitors. Dev Biol. 2012;363:373–387. doi: 10.1016/j.ydbio.2012.01.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Musser MA, Correa H, Southard-Smith EM. Enteric neuron imbalance and proximal dysmotility in ganglionated intestine of the Hirschsprung mouse model. Cell Mol Gastroenterol Hepatol. 2015;1:87–101. doi: 10.1016/j.jcmgh.2014.08.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  111. Musser MA, Southard-Smith EM. Balancing on the crest - Evidence for disruption of the enteric ganglia via inappropriate lineage segregation and consequences for gastrointestinal function. Dev Biol. 2013;382:356–364. doi: 10.1016/j.ydbio.2013.01.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Nagashimada M, Ohta H, Li C, Nakao K, Uesaka T, Brunet JF, Amiel J, Trochet D, Wakayama T, Enomoto H. Autonomic neurocristopathy-associated mutations in PHOX2B dysregulate Sox10 expression. J Clin Invest. 2012;122:3145–3158. doi: 10.1172/JCI63401. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Nagy N, Goldstein AM. Endothelin-3 regulates neural crest cell proliferation and differentiation in the hindgut enteric nervous system. Dev Biol. 2006;293:203–217. doi: 10.1016/j.ydbio.2006.01.032. [DOI] [PubMed] [Google Scholar]
  114. Nagy N, Mwizerwa O, Yaniv K, Carmel L, Pieretti-Vanmarcke R, Weinstein BM, Goldstein AM. Endothelial cells promote migration and proliferation of enteric neural crest cells via beta1 integrin signaling. Dev Biol. 2009;330:263–272. doi: 10.1016/j.ydbio.2009.03.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Ngan ES, Garcia-Barcelo MM, Yip BH, Poon HC, Lau ST, Kwok CK, Sat E, Sham MH, Wong KK, Wainwright BJ, Cherny SS, Hui CC, Sham PC, Lui VC, Tam PK. Hedgehog/Notch-induced premature gliogenesis represents a new disease mechanism for Hirschsprung disease in mice and humans. J Clin Invest. 2011;121:3467–3478. doi: 10.1172/JCI43737. [DOI] [PMC free article] [PubMed] [Google Scholar]
  116. Nie X, Wang Q, Jiao K. Dicer activity in neural crest cells is essential for craniofacial organogenesis and pharyngeal arch artery morphogenesis. Mech Dev. 2011;128:200–207. doi: 10.1016/j.mod.2010.12.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Niederreither K, Vermot J, Le Roux I, Schuhbaur B, Chambon P, Dolle P. The regional pattern of retinoic acid synthesis by RALDH2 is essential for the development of posterior pharyngeal arches and the enteric nervous system. Development. 2003;130:2525–2534. doi: 10.1242/dev.00463. [DOI] [PubMed] [Google Scholar]
  118. Nishiyama C, Uesaka T, Manabe T, Yonekura Y, Nagasawa T, Newgreen DF, Young HM, Enomoto H. Trans-mesenteric neural crest cells are the principal source of the colonic enteric nervous system. Nat Neurosci. 2012;15:1211–1218. doi: 10.1038/nn.3184. [DOI] [PubMed] [Google Scholar]
  119. Obermayr F, Hotta R, Enomoto H, Young HM. Development and developmental disorders of the enteric nervous system. Nat Rev Gastroenterol Hepatol. 2013;10:43–57. doi: 10.1038/nrgastro.2012.234. [DOI] [PubMed] [Google Scholar]
  120. Okamura Y, Saga Y. Notch signaling is required for the maintenance of enteric neural crest progenitors. Development. 2008;135:3555–3565. doi: 10.1242/dev.022319. [DOI] [PubMed] [Google Scholar]
  121. Owens SE, Broman KW, Wiltshire T, Elmore JB, Bradley KM, Smith JR, Southard-Smith EM. Genome-wide linkage identifies novel modifier loci of aganglionosis in the Sox10Dom model of Hirschsprung disease. Hum. Mol. Genet. 2005;14(11):1549–1558. doi: 10.1093/hmg/ddi163. [DOI] [PubMed] [Google Scholar]
  122. Paratore C, Eichenberger C, Suter U, Sommer L. Sox10 haploinsufficiency affects maintenance of progenitor cells in a mouse model of Hirschsprung disease. Hum Mol Genet. 2002;11:3075–3085. doi: 10.1093/hmg/11.24.3075. [DOI] [PubMed] [Google Scholar]
  123. Paratore C, Goerich DE, Suter U, Wegner M, Sommer L. Survival and glial fate acquisition of neural crest cells are regulated by an interplay between the transcription factor Sox10 and extrinsic combinatorial signaling. Development. 2001;128:3949–3961. doi: 10.1242/dev.128.20.3949. [DOI] [PubMed] [Google Scholar]
  124. Pattyn A, Morin X, Cremer H, Goridis C, Brunet JF. The homeobox gene Phox2b is essential for the development of autonomic neural crest derivatives. Nature. 1999;399:366–370. doi: 10.1038/20700. [DOI] [PubMed] [Google Scholar]
  125. Pham TD, Gershon MD, Rothman TP. Time of origin of neurons in the murine enteric nervous system: sequence in relation to phenotype. J Comp Neurol. 1991;314:789–798. doi: 10.1002/cne.903140411. [DOI] [PubMed] [Google Scholar]
  126. Pichel JG, Shen L, Sheng HZ, Granholm AC, Drago J, Grinberg A, Lee EJ, Huang SP, Saarma M, Hoffer BJ, Sariola H, Westphal H. Defects in enteric innervation and kidney development in mice lacking GDNF. Nature. 1996a;382:73–76. doi: 10.1038/382073a0. [DOI] [PubMed] [Google Scholar]
  127. Pichel JG, Shen L, Sheng HZ, Granholm AC, Drago J, Grinberg A, Lee EJ, Huang SP, Saarma M, Hoffer BJ, Sariola H, Westphal H. GDNF is required for kidney development and enteric innervation. Cold Spring Harb Symp Quant Biol. 1996b;61:445–457. [PubMed] [Google Scholar]
  128. Pingault V, Bondurand N, Kuhlbrodt K, Goerich DE, Prehu MO, Puliti A, Herbarth B, Hermans-Borgmeyer I, Legius E, Matthijs G, Amiel J, Lyonnet S, Ceccherini I, Romeo G, Smith JC, Read AP, Wegner M, Goossens M. SOX10 mutations in patients with Waardenburg-Hirschsprung disease. Nat Genet. 1998;18:171–173. doi: 10.1038/ng0298-171. [DOI] [PubMed] [Google Scholar]
  129. Pingault V, Ente D, Dastot-Le Moal F, Goossens M, Marlin S, Bondurand N. Review and update of mutations causing Waardenburg syndrome. Hum Mutat. 2010;31:391–406. doi: 10.1002/humu.21211. [DOI] [PubMed] [Google Scholar]
  130. Pritchard CA, Bolin L, Slattery R, Murray R, McMahon M. Post-natal lethality and neurological and gastrointestinal defects in mice with targeted disruption of the A-Raf protein kinase gene. Curr Biol. 1996;6:614–617. doi: 10.1016/s0960-9822(02)00548-1. [DOI] [PubMed] [Google Scholar]
  131. Puffenberger EG, Kauffman ER, Bolk S, Matise TC, Washington SS, Angrist M, Weissenbach J, Garver KL, Mascari M, Ladda R, et al. Identity-by-descent and association mapping of a recessive gene for Hirschsprung disease on human chromosome 13q22. Hum Mol Genet. 1994;3:1217–1225. doi: 10.1093/hmg/3.8.1217. [DOI] [PubMed] [Google Scholar]
  132. Puig I, Champeval D, De Santa Barbara P, Jaubert F, Lyonnet S, Larue L. Deletion of Pten in the mouse enteric nervous system induces ganglioneuromatosis and mimics intestinal pseudoobstruction. J Clin Invest. 2009;119:3586–3596. doi: 10.1172/JCI39929. [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. Qiu M, Bulfone A, Martinez S, Meneses JJ, Shimamura K, Pedersen RA, Rubenstein JL. Null mutation of Dlx-2 results in abnormal morphogenesis of proximal first and second branchial arch derivatives and abnormal differentiation in the forebrain. Genes Dev. 1995;9:2523–2538. doi: 10.1101/gad.9.20.2523. [DOI] [PubMed] [Google Scholar]
  134. Ramalho-Santos M, Melton DA, McMahon AP. Hedgehog signals regulate multiple aspects of gastrointestinal development. Development. 2000;127:2763–2772. doi: 10.1242/dev.127.12.2763. [DOI] [PubMed] [Google Scholar]
  135. Riethmacher D, Sonnenberg-Riethmacher E, Brinkmann V, Yamaai T, Lewin GR, Birchmeier C. Severe neuropathies in mice with targeted mutations in the ErbB3 receptor. Nature. 1997;389:725–730. doi: 10.1038/39593. [DOI] [PubMed] [Google Scholar]
  136. Roberts RR, Bornstein JC, Bergner AJ, Young HM. Disturbances of colonic motility in mouse models of Hirschsprung’s disease. Am J Physiol Gastrointest Liver Physiol. 2008;294:G996–G1008. doi: 10.1152/ajpgi.00558.2007. [DOI] [PubMed] [Google Scholar]
  137. Romeo G, Ronchetto P, Luo Y, Barone V, Seri M, Ceccherini I, Pasini B, Bocciardi R, Lerone M, Kaariainen H, et al. Point mutations affecting the tyrosine kinase domain of the RET proto-oncogene in Hirschsprung’s disease. Nature. 1994;367:377–378. doi: 10.1038/367377a0. [DOI] [PubMed] [Google Scholar]
  138. Rossi J, Herzig KH, Voikar V, Hiltunen PH, Segerstrale M, Airaksinen MS. Alimentary tract innervation deficits and dysfunction in mice lacking GDNF family receptor alpha2. J Clin Invest. 2003;112:707–716. doi: 10.1172/JCI17995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  139. Rossi J, Luukko K, Poteryaev D, Laurikainen A, Sun YF, Laakso T, Eerikainen S, Tuominen R, Lakso M, Rauvala H, Arumae U, Pasternack M, Saarma M, Airaksinen MS. Retarded growth and deficits in the enteric and parasympathetic nervous system in mice lacking GFR alpha2, a functional neurturin receptor. Neuron. 1999;22:243–252. doi: 10.1016/s0896-6273(00)81086-7. [DOI] [PubMed] [Google Scholar]
  140. Rothman TP, Gershon MD. Regionally defective colonization of the terminal bowel by the precursors of enteric neurons in lethal spotted mutant mice. Neuroscience. 1984;12:1293–1311. doi: 10.1016/0306-4522(84)90022-8. [DOI] [PubMed] [Google Scholar]
  141. Saldana-Caboverde A, Perera EM, Watkins-Chow DE, Hansen NF, Vemulapalli M, Mullikin JC, Program NCS, Pavan WJ, Kos L. The transcription factors Ets1 and Sox10 interact during murine melanocyte development. Dev Biol. 2015;407:300–312. doi: 10.1016/j.ydbio.2015.04.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Sanchez MP, Silos-Santiago I, Frisen J, He B, Lira SA, Barbacid M. Renal agenesis and the absence of enteric neurons in mice lacking GDNF. Nature. 1996;382:70–73. doi: 10.1038/382070a0. [DOI] [PubMed] [Google Scholar]
  143. Sanchez-Ferras O, Bernas G, Farnos O, Toure AM, Souchkova O, Pilon N. A direct role for murine Cdx proteins in the trunk neural crest gene regulatory network. Development. 2016;143:1363–1374. doi: 10.1242/dev.132159. [DOI] [PubMed] [Google Scholar]
  144. Sasselli V, Boesmans W, Vanden Berghe P, Tissir F, Goffinet AM, Pachnis V. Planar cell polarity genes control the connectivity of enteric neurons. J Clin Invest. 2013;123:1763–1772. doi: 10.1172/JCI66759. [DOI] [PMC free article] [PubMed] [Google Scholar]
  145. Sato Y, Heuckeroth RO. Retinoic acid regulates murine enteric nervous system precursor proliferation, enhances neuronal precursor differentiation, and reduces neurite growth in vitro. Dev Biol. 2008;320:185–198. doi: 10.1016/j.ydbio.2008.05.524. [DOI] [PMC free article] [PubMed] [Google Scholar]
  146. Schill EM, Lake JI, Tusheva OA, Nagy N, Bery SK, Foster L, Avetisyan M, Johnson SL, Stenson WF, Goldstein AM, Heuckeroth RO. Ibuprofen slows migration and inhibits bowel colonization by enteric nervous system precursors in zebrafish, chick and mouse. Dev Biol. 2016;409:473–488. doi: 10.1016/j.ydbio.2015.09.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  147. Schuchardt A, D’Agati V, Larsson-Blomberg L, Costantini F, Pachnis V. Defects in the kidney and enteric nervous system of mice lacking the tyrosine kinase receptor Ret. Nature. 1994;367:380–383. doi: 10.1038/367380a0. [DOI] [PubMed] [Google Scholar]
  148. Sharan A, Zhu H, Xie H, Li H, Tang J, Tang W, Zhang H, Xia Y. Down-regulation of miR-206 is associated with Hirschsprung disease and suppresses cell migration and proliferation in cell models. Sci Rep. 2015;5:9302. doi: 10.1038/srep09302. [DOI] [PMC free article] [PubMed] [Google Scholar]
  149. Shen L, Pichel JG, Mayeli T, Sariola H, Lu B, Westphal H. Gdnf haploinsufficiency causes Hirschsprung-like intestinal obstruction and early-onset lethality in mice. Am J Hum Genet. 2002;70:435–447. doi: 10.1086/338712. [DOI] [PMC free article] [PubMed] [Google Scholar]
  150. Shirasawa S, Yunker AM, Roth KA, Brown GA, Horning S, Korsmeyer SJ. Enx (Hox11L1)-deficient mice develop myenteric neuronal hyperplasia and megacolon. Nat Med. 1997;3:646–650. doi: 10.1038/nm0697-646. [DOI] [PubMed] [Google Scholar]
  151. Soret R, Mennetrey M, Bergeron KF, Dariel A, Neunlist M, Grunder F, Faure C, Silversides DW, Pilon N, Ente-Hirsch Study G. A collagen VI-dependent pathogenic mechanism for Hirschsprung’s disease. J Clin Invest. 2015;125:4483–4496. doi: 10.1172/JCI83178. [DOI] [PMC free article] [PubMed] [Google Scholar]
  152. Southard-Smith EM, Kos L, Pavan WJ. Sox10 mutation disrupts neural crest development in Dom Hirschsprung mouse model. Nat Genet. 1998;18:60–64. doi: 10.1038/ng0198-60. [DOI] [PubMed] [Google Scholar]
  153. Spouge D, Baird PA. Hirschsprung disease in a large birth cohort. Teratology. 1985;32:171–177. doi: 10.1002/tera.1420320204. [DOI] [PubMed] [Google Scholar]
  154. Sribudiani Y, Metzger M, Osinga J, Rey A, Burns AJ, Thapar N, Hofstra RM. Variants in RET associated with Hirschsprung’s disease affect binding of transcription factors and gene expression. Gastroenterology. 2011;140:572–582. e572. doi: 10.1053/j.gastro.2010.10.044. [DOI] [PubMed] [Google Scholar]
  155. Stanchina L, Baral V, Robert F, Pingault V, Lemort N, Pachnis V, Goossens M, Bondurand N. Interactions between Sox10, Edn3 and Ednrb during enteric nervous system and melanocyte development. Dev Biol. 2006;295:232–249. doi: 10.1016/j.ydbio.2006.03.031. [DOI] [PubMed] [Google Scholar]
  156. Stanchina L, Van de Putte T, Goossens M, Huylebroeck D, Bondurand N. Genetic interaction between Sox10 and Zfhx1b during enteric nervous system development. Dev Biol. 2010;341:416–428. doi: 10.1016/j.ydbio.2010.02.036. [DOI] [PubMed] [Google Scholar]
  157. Taketomi T, Yoshiga D, Taniguchi K, Kobayashi T, Nonami A, Kato R, Sasaki M, Sasaki A, Ishibashi H, Moriyama M, Nakamura K, Nishimura J, Yoshimura A. Loss of mammalian Sprouty2 leads to enteric neuronal hyperplasia and esophageal achalasia. Nat Neurosci. 2005;8:855–857. doi: 10.1038/nn1485. [DOI] [PubMed] [Google Scholar]
  158. Tang W, Li H, Tang J, Wu W, Qin J, Lei H, Cai P, Huo W, Li B, Rehan V, Xu X, Geng Q, Zhang H, Xia Y. Specific serum microRNA profile in the molecular diagnosis of Hirschsprung’s disease. J Cell Mol Med. 2014;18:1580–1587. doi: 10.1111/jcmm.12348. [DOI] [PMC free article] [PubMed] [Google Scholar]
  159. Teng L, Mundell NA, Frist AY, Wang Q, Labosky PA. Requirement for Foxd3 in the maintenance of neural crest progenitors. Development. 2008;135:1615–1624. doi: 10.1242/dev.012179. [DOI] [PMC free article] [PubMed] [Google Scholar]
  160. Threadgill DW, Hunter KW, Williams RW. Genetic dissection of complex and quantitative traits: from fantasy to reality via a community effort. Mamm Genome. 2002;13:175–178. doi: 10.1007/s00335-001-4001-Y. [DOI] [PubMed] [Google Scholar]
  161. Uesaka T, Enomoto H. Neural precursor death is central to the pathogenesis of intestinal aganglionosis in Ret hypomorphic mice. J Neurosci. 2010;30:5211–5218. doi: 10.1523/JNEUROSCI.6244-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  162. Uesaka T, Jain S, Yonemura S, Uchiyama Y, Milbrandt J, Enomoto H. Conditional ablation of GFRalpha1 in postmigratory enteric neurons triggers unconventional neuronal death in the colon and causes a Hirschsprung’s disease phenotype. Development. 2007;134:2171–2181. doi: 10.1242/dev.001388. [DOI] [PubMed] [Google Scholar]
  163. Uesaka T, Nagashimada M, Enomoto H. Neuronal Differentiation in Schwann Cell Lineage Underlies Postnatal Neurogenesis in the Enteric Nervous System. J Neurosci. 2015;35:9879–9888. doi: 10.1523/JNEUROSCI.1239-15.2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  164. Uesaka T, Nagashimada M, Yonemura S, Enomoto H. Diminished Ret expression compromises neuronal survival in the colon and causes intestinal aganglionosis in mice. J Clin Invest. 2008;118:1890–1898. doi: 10.1172/JCI34425. [DOI] [PMC free article] [PubMed] [Google Scholar]
  165. Uesaka T, Young HM, Pachnis V, Enomoto H. Development of the intrinsic and extrinsic innervation of the gut. Dev Biol. 2016 doi: 10.1016/j.ydbio.2016.04.016. [DOI] [PubMed] [Google Scholar]
  166. Van de dePutte T, Francis A, Nelles L, van Grunsven LA, Huylebroeck D. Neural crest-specific removal of Zfhx1b in mouse leads to a wide range of neurocristopathies reminiscent of Mowat-Wilson syndrome. Hum Mol Genet. 2007;16:1423–1436. doi: 10.1093/hmg/ddm093. [DOI] [PubMed] [Google Scholar]
  167. Van de Putte T, Maruhashi M, Francis A, Nelles L, Kondoh H, Huylebroeck D, Higashi Y. Mice lacking ZFHX1B, the gene that codes for Smad-interacting protein-1, reveal a role for multiple neural crest cell defects in the etiology of Hirschsprung disease-mental retardation syndrome. Am J Hum Genet. 2003;72:465–470. doi: 10.1086/346092. [DOI] [PMC free article] [PubMed] [Google Scholar]
  168. Viader A, Wright-Jin EC, Vohra BP, Heuckeroth RO, Milbrandt J. Differential regional and subtype-specific vulnerability of enteric neurons to mitochondrial dysfunction. PLoS One. 2011;6:e27727. doi: 10.1371/journal.pone.0027727. [DOI] [PMC free article] [PubMed] [Google Scholar]
  169. Wahlbuhl M, Reiprich S, Vogl MR, Bosl MR, Wegner M. Transcription factor Sox10 orchestrates activity of a neural crest-specific enhancer in the vicinity of its gene. Nucleic Acids Res. 2012;40:88–101. doi: 10.1093/nar/gkr734. [DOI] [PMC free article] [PubMed] [Google Scholar]
  170. Wakamatsu N, Yamada Y, Yamada K, Ono T, Nomura N, Taniguchi H, Kitoh H, Mutoh N, Yamanaka T, Mushiake K, Kato K, Sonta S, Nagaya M. Mutations in SIP1, encoding Smad interacting protein-1, cause a form of Hirschsprung disease. Nat Genet. 2001;27:369–370. doi: 10.1038/86860. [DOI] [PubMed] [Google Scholar]
  171. Wallace AS, Barlow AJ, Navaratne L, Delalande JM, Tauszig-Delamasure S, Corset V, Thapar N, Burns AJ. Inhibition of cell death results in hyperganglionosis: implications for enteric nervous system development. Neurogastroenterol Motil. 2009;21:768–e749. doi: 10.1111/j.1365-2982.2009.01309.x. [DOI] [PubMed] [Google Scholar]
  172. Wallace AS, Tan MX, Schachner M, Anderson RB. L1cam acts as a modifier gene for members of the endothelin signalling pathway during enteric nervous system development. Neurogastroenterol Motil. 2011;23:e510–e522. doi: 10.1111/j.1365-2982.2011.01692.x. [DOI] [PubMed] [Google Scholar]
  173. Walters LC, Cantrell VA, Weller KP, Mosher JT, Southard-Smith EM. Genetic background impacts developmental potential of enteric neural crest-derived progenitors in the Sox10Dom model of Hirschsprung disease. Hum Mol Genet. 2010;19:4353–4372. doi: 10.1093/hmg/ddq357. [DOI] [PMC free article] [PubMed] [Google Scholar]
  174. Wang H, Hughes I, Planer W, Parsadanian A, Grider JR, Vohra BP, Keller-Peck C, Heuckeroth RO. The timing and location of glial cell line-derived neurotrophic factor expression determine enteric nervous system structure and function. J Neurosci. 2010;30:1523–1538. doi: 10.1523/JNEUROSCI.3861-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  175. Wang X, Chan AK, Sham MH, Burns AJ, Chan WY. Analysis of the sacral neural crest cell contribution to the hindgut enteric nervous system in the mouse embryo. Gastroenterology. 2011;141:992–1002. e1001–e1006. doi: 10.1053/j.gastro.2011.06.002. [DOI] [PubMed] [Google Scholar]
  176. Watanabe Y, Broders-Bondon F, Baral V, Paul-Gilloteaux P, Pingault V, Dufour S, Bondurand N. Sox10 and Itgb1 interaction in enteric neural crest cell migration. Dev Biol. 2013;379:92–106. doi: 10.1016/j.ydbio.2013.04.013. [DOI] [PubMed] [Google Scholar]
  177. Webster W. Embryogenesis of the enteric ganglia in normal mice and in mice that develop congenital aganglionic megacolon. J Embryol Exp Morphol. 1973;30:573–585. [PubMed] [Google Scholar]
  178. Wright-Jin EC, Grider JR, Duester G, Heuckeroth RO. Retinaldehyde dehydrogenase enzymes regulate colon enteric nervous system structure and function. Dev Biol. 2013;381:28–37. doi: 10.1016/j.ydbio.2013.06.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  179. Wu JJ, Chen JX, Rothman TP, Gershon MD. Inhibition of in vitro enteric neuronal development by endothelin-3: mediation by endothelin B receptors. Development. 1999;126:1161–1173. doi: 10.1242/dev.126.6.1161. [DOI] [PubMed] [Google Scholar]
  180. Yamada T, Ohtani S, Sakurai T, Tsuji T, Kunieda T, Yanagisawa M. Reduced expression of the endothelin receptor type B gene in piebald mice caused by insertion of a retroposon-like element in intron 1. J Biol Chem. 2006;281:10799–10807. doi: 10.1074/jbc.M512618200. [DOI] [PubMed] [Google Scholar]
  181. Yan H, Bergner AJ, Enomoto H, Milbrandt J, Newgreen DF, Young HM. Neural cells in the esophagus respond to glial cell line-derived neurotrophic factor and neurturin, and are RET-dependent. Dev Biol. 2004;272:118–133. doi: 10.1016/j.ydbio.2004.04.025. [DOI] [PubMed] [Google Scholar]
  182. Yanagisawa H, Yanagisawa M, Kapur RP, Richardson JA, Williams SC, Clouthier DE, de Wit D, Emoto N, Hammer RE. Dual genetic pathways of endothelin-mediated intercellular signaling revealed by targeted disruption of endothelin converting enzyme-1 gene. Development. 1998;125:825–836. doi: 10.1242/dev.125.5.825. [DOI] [PubMed] [Google Scholar]
  183. Yang JT, Liu CZ, Villavicencio EH, Yoon JW, Walterhouse D, Iannaccone PM. Expression of human GLI in mice results in failure to thrive, early death, and patchy Hirschsprung-like gastrointestinal dilatation. Mol Med. 1997;3:826–835. [PMC free article] [PubMed] [Google Scholar]
  184. Young HM. Physiology of the Gastrointestinal Tract. 5th. London, UK: Elsevier; 2012. [Google Scholar]
  185. Young HM, Bergner AJ, Simpson MJ, McKeown SJ, Hao MM, Anderson CR, Enomoto H. Colonizing while migrating: how do individual enteric neural crest cells behave? BMC Biol. 2014;12:23. doi: 10.1186/1741-7007-12-23. [DOI] [PMC free article] [PubMed] [Google Scholar]
  186. Young HM, Hearn CJ, Ciampoli D, Southwell BR, Brunet JF, Newgreen DF. A single rostrocaudal colonization of the rodent intestine by enteric neuron precursors is revealed by the expression of Phox2b, Ret, and p75 and by explants grown under the kidney capsule or in organ culture. Dev Biol. 1998;202:67–84. doi: 10.1006/dbio.1998.8987. [DOI] [PubMed] [Google Scholar]
  187. Zehir A, Hua LL, Maska EL, Morikawa Y, Cserjesi P. Dicer is required for survival of differentiating neural crest cells. Dev Biol. 2010;340:459–467. doi: 10.1016/j.ydbio.2010.01.039. [DOI] [PMC free article] [PubMed] [Google Scholar]
  188. Zhang B, Jain S, Song H, Fu M, Heuckeroth RO, Erlich JM, Jay PY, Milbrandt J. Mice lacking sister chromatid cohesion protein PDS5B exhibit developmental abnormalities reminiscent of Cornelia de Lange syndrome. Development. 2007;134:3191–3201. doi: 10.1242/dev.005884. [DOI] [PubMed] [Google Scholar]
  189. Zhang Y, Kim TH, Niswander L. Phactr4 regulates directional migration of enteric neural crest through PP1, integrin signaling, and cofilin activity. Genes Dev. 2012;26:69–81. doi: 10.1101/gad.179283.111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  190. Zhang Y, Niswander L. Phactr4: a new integrin modulator required for directional migration of enteric neural crest cells. Cell Adh Migr. 2012;6:419–423. doi: 10.4161/cam.21266. [DOI] [PMC free article] [PubMed] [Google Scholar]
  191. Zhang Y, Niswander L. Zic2 is required for enteric nervous system development and neurite outgrowth: a mouse model of enteric hyperplasia and dysplasia. Neurogastroenterol Motil. 2013;25:538–541. doi: 10.1111/nmo.12101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  192. Zhou R, Niwa S, Homma N, Takei Y, Hirokawa N. KIF26A is an unconventional kinesin and regulates GDNF-Ret signaling in enteric neuronal development. Cell. 2009;139:802–813. doi: 10.1016/j.cell.2009.10.023. [DOI] [PubMed] [Google Scholar]
  193. Zhu D, Xie H, Li H, Cai P, Zhu H, Xu C, Chen P, Sharan A, Xia Y, Tang W. Nidogen-1 is a common target of microRNAs MiR-192/215 in the pathogenesis of Hirschsprung’s disease. J Neurochem. 2015;134:39–46. doi: 10.1111/jnc.13118. [DOI] [PubMed] [Google Scholar]
  194. Zhu L, Lee HO, Jordan CS, Cantrell VA, Southard-Smith EM, Shin MK. Spatiotemporal regulation of endothelin receptor-B by SOX10 in neural crest-derived enteric neuron precursors. Nat Genet. 2004;36:732–737. doi: 10.1038/ng1371. [DOI] [PubMed] [Google Scholar]
  195. Zimmer J, Puri P. Knockout mouse models of Hirschsprung’s disease. Pediatr Surg Int. 2015;31:787–794. doi: 10.1007/s00383-015-3747-3. [DOI] [PubMed] [Google Scholar]

RESOURCES