Developmental Neuroendocrinology: Susan Wray Seth Blackshaw Editors
Developmental Neuroendocrinology: Susan Wray Seth Blackshaw Editors
Susan Wray
Seth Blackshaw Editors
Developmental
Neuroendocrinology
Masterclass in Neuroendocrinology
Volume 9
Series Editors
John A. Russell
Centre for Discovery Brain Sciences, Edinburgh University Medical School,
CMVM, Edinburgh, UK
William E. Armstrong
Department of Anatomy and Neurobiology, The University of Tennessee Health
Science Center, Memphis, TN, USA
Masterclass in Neuroendocrinology is a co-publication of the INF (International
Neuroendocrine Federation) that aims to illustrate the highest standards and promote
the use of the latest technologies in basic and clinical research, while also providing
inspiration for further exploration into the exciting field of neuroendocrinology. It is
intended for established researchers, trainees and students alike.
Each book
- is edited by leading experts in the field
- is written by a team of internationally respected researchers
- includes assessments of different experimental approaches, both in vivo and in
vitro, and of how the resulting data are interpreted.
Developmental
Neuroendocrinology
Editors
Susan Wray Seth Blackshaw
Cellular and Developmental Neurobiology Department of Neuroscience
NINDS, NIH School of Medicine, Johns Hopkins University
Bethesda, USA Baltimore, MD, USA
This Springer imprint is published by the registered company Springer Nature Switzerland AG.
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Series Preface
This series began publication as a joint venture between the International Neuroen-
docrine Federation and Wiley‐Blackwell and now is continuing with Springer-
Nature as publisher for the federation. The broad aim of the series is to provide
established researchers, trainees, and students with authoritative up‐to‐date accounts
of the present state of knowledge and prospects for the future across a range of topics
in the burgeoning field of neuroendocrinology. The series is aimed at a wide
audience as neuroendocrinology integrates neuroscience and endocrinology. We
define neuroendocrinology as the study of the control of endocrine function by the
brain and the actions of hormones on the brain. It encompasses the study of normal
and abnormal function and the developmental origins of disease. It includes the
study of the neural networks in the brain that regulate and form neuroendocrine
systems. It also includes the study of behaviors and mental states that are influenced
or regulated by hormones. It necessarily includes the understanding and study of
peripheral physiological systems that are regulated by neuroendocrine mechanisms.
Clearly, neuroendocrinology embraces many current issues of concern to human
health and well‐being, but research on these issues necessitates reductionist animal
models.
Contemporary research in neuroendocrinology involves the use of a wide range
of techniques and technologies, from the subcellular to systems at the whole-
organism level. A particular aim of the series is to provide expert advice and
discussion about experimental or study protocols in research in neuroendocrinology
and to further advance the field by giving information and advice about novel
techniques, technologies, and interdisciplinary approaches.
To achieve our aims, each book is on a particular theme in neuroendocrinology,
and for each book we have recruited a pair of editors, expert in the field, and they
have engaged an international team of experts to contribute chapters in their individ-
ual areas of expertise. Their mission was to give an update of knowledge and recent
discoveries and to discuss new approaches, “gold‐standard” protocols, translational
possibilities, and future prospects. Authors were asked to write for a wide audience,
to minimize references, and to consider the use of video clips and explanatory text
boxes; each chapter is peer reviewed and has a Glossary and a detailed Index. We
have been guided by an Advisory Editorial Board.
v
vi Series Preface
The MasterClass Series is open‐ended; books in the series published to date are
Neurophysiology of Neuroendocrine Neurons (2014, ed. WE Armstrong & JG
Tasker); Neuroendocrinology of Stress (2015, ed. JA Russell & MJ Shipston);
Molecular Neuroendocrinology: From Genome to Physiology (2016, ed. D Murphy
& H Gainer); Computational Neuroendocrinology (2016, ed. DJ Macgregor & G
Leng); Neuroendocrinology of Appetite (2016; ed. SL Dickson & JG Mercer); The
GnRH Neuron and its Control (2018; ed. AE Herbison & TM Plant); Model Animals
in Neuroendocrinology (2019, ed. M Ludwig and G Levkowitz); and Neurosecre-
tion: Secretory Mechanisms (In Press, ed. J Lemos and G Dayanithi), #8 in the
Series, and will be the first to be published by Springer Nature; books in preparation
include Glial-Neuronal Signaling in Neuroendocrine Systems and Neuroendocrine
Clocks and Calendars.
Feedback and suggestions are welcome.
vii
viii Volume Preface
ix
x Contents
Glossary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 447
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 459
Part I
Molecular Specification of Hypothalamic/
Pituitary Cells
Development of the Neuroendocrine
Hypothalamus 1
Marysia Placzek, Travis Fu, and Matthew Towers
Abstract
The neuroendocrine hypothalamus consists of neurosecretory neurons and
specialized glia of the median eminence and posterior pituitary. Neurons and
glia act together to support and regulate reciprocal brain–body communications
that enable the hypothalamus to anticipate and adapt to changing physiological
conditions. Here we summarize recent studies in model organisms that show how
an embryonic program of hypothalamic development unfolds in a manner that
underlies adult function. Fate-mapping studies in chick show that much of the
basal hypothalamus is built from a multipotent Fgf10(+) progenitor population
whose mode of growth in four dimensions suggests a new model for hypotha-
lamic development that we term the ‘Anisotropic growth’ model. Conditional
genetic analyses and pharmacological interventions in mouse, zebrafish, and
chick suggest that a conserved molecular mechanism may mediate anisotropic
growth and establish the basal hypothalamus. A handful of conserved signalling
factors and transcription factors (TFs) direct the progressive development of
Fgf10(+) progenitors to hypothalamic progenitors, hypothalamic neurons, and
then infundibular progenitors of the median eminence and posterior pituitary in a
program where space and time are intrinsically linked. These studies reveal
additionally that Fgf10(+) cell populations are maintained throughout life and
are vital to the long-term building, maintenance, and function of the adult
hypothalamo–pituitary axis. We speculate on whether the embryonic Fgf10(+)
progenitor population harbours a hypothalamic stem cell and discuss the
implications for the ability of the hypothalamus to adapt to changing physiologi-
cal conditions on different time scales through life.
Keywords
Fgf10 · Shh · Anisotropic growth · Infundibulum · Median eminence · Posterior
pituitary
List of Abbreviations
3V 3rd ventricle
AP Anterior Pituitary (aka Adenohypophysis)
ARC Arcuate nucleus
bHLH Basic helix-loop-helix
BMP Bone morphogenetic protein
CRH Corticotropin releasing hormone
DA Dopamine
DMN Dorsomedial nucleus
Fgf Fibroblast growth factor
GHRH Growth hormone releasing hormone
HD Homeodomain
MB Mammillary Body
ME Median Eminence
OXT Oxytocin
PeVN Periventricular nucleus
PP Posterior pituitary (aka neurohypophysis)
PVN Paraventricular nucleus
RDVM Rostral diencephalic ventral midline
Rx Retina and anterior neural fold homeobox
SCN Suprachiasmatic nucleus
Shh Sonic hedgehog
SON Supraoptic nucleus
SST Somatostatin
TMN Tubero-mammillary nucleus
TRH Thyroid releasing hormone
VMN Ventromedial nucleus
VZ Ventricular zone
Fig.1.1 The adult neuroendocrine hypothalamus. (a) Schematic side view of adult brain. (b) High-
power view of boxed region in (a) to show hypothalamus and approximate position of hypotha-
lamic nuclei relative to optic stalk (OS). (c, d) Transverse views through planes indicated in (b): at
level of anterior hypothalamus (c) or ME (d) to show relative positions of neuroendocrine nuclei
and glia. Different tanycyte subsets are indicated by colour: β2 green: β1 yellow; α2 purple, α1 blue.
(e) High power view of boxed region in (b) to show approximate positions of ME, PP and AP. In the
ME, parvocellular neuronal endfeet intimately contact tanycytes and portal capillaries. In the PP,
magnocellular neuronal endfeet contact pituicytes and capillaries
over different time scales: for instance, short-term changes in stress/energy balance,
or long-term seasonal or lifecourse changes in energy balance and reproduction.
Anatomically and functionally, the hypothalamus is complex. In much of the
central nervous system, such as the cortex, cerebellum, hindbrain or spinal cord,
functionally similar neurons are arranged in columns or layers. By contrast, the
hypothalamus is composed of nuclei, or areas, arranged in a patchwork around the
third ventricle. Each nucleus/area is located in a stereotyped position in the adult
(Fig. 1.1a, b) and harbours functionally diverse neurons. Neuroendocrine neurons
6 M. Placzek et al.
Fig. 1.2 Position and growth of chick bHyp cells. (a, b) Schematics of 10-somite stage chick
embryos: ventral (a) and sagittal (b) views. Fgf10(+) bHyp progenitors (red) are nested within
diencephalic progenitors (yellow: note in mouse these express Foxd1), abut Foxg1(+) telencephalic
progenitor cells (green) anteriorly and lie above the prechordal mesendoderm (pm). Arrows indicate
subsequent bidirectional growth (c) Side view, 25 somite embryo. bHyp progenitors give rise to
descendants along the entire basal hypothalamus (red indicates all bHyp-derived cells: note Fgf10 is
maintained only centrally (between arrows). Arrows indicate bidirectional growth
Fig. 1.3 Three-dimensional sequential anisotropic growth from bHyp cells. Schematic sagittal
views of chick embryo (10–40 somites). (a) bHyp cells (red) abut the telencephalon (green) in the
10-somite embryo. (b) By 12 somites, bHyp cells begin to generate anterior progenitors (orange).
(c) By 25 somites, mammillary progenitors are generated (blue): these extend posteriorly from
bHyp cells that are now central (red). (d) Finally, infundibular glial cells are generated and grow
ventrally (arrows). Dotted circle indicates optic stalk (os). Basal hypothalamus shown relative to
underlying tissues: prechordal mesendoderm (a); developing (b) or definitive (c, d) Rathke’s pouch
tangential growth of anterior and mammillary progenitor cells from bHyp cells
means that the adult neurons to which they give rise, including neuroendocrine
neurons, may ultimately lie distant from the sites of their progenitor cell
specification.
10 M. Placzek et al.
(A) The columnar model (Swanson 1992) derives from classical studies in the
rat. It suggests that the hypothalamus is a diencephalic-derived structure
whose neurons/nuclei fall into different areas along the rostro–caudal
axis: preoptic/anterior, tuberal, and mammillary, reflecting distinct
diencephalic progenitor subsets arranged in columns along the anterior–
posterior (future rostro–caudal) axis. In this model, neuroendocrine
neurons of the PVN, PeVN, SON and SCN derive from preoptic/anterior
progenitors whereas those of the Arc derive from tuberal progenitors.
Recent single cell RNA-seq studies of wild-type and mutant mice (Kim
et al. 2019) support a version of the columnar model.
(B) The revised prosomeric model (Puelles et al. 2012, 2013), based on gene
expression profiling and genetic inducible fate-mapping studies in mouse
and chick (Sánchez-Arrones et al. 2009; Alvarez-Bolado et al. 2012)
(continued)
1 Development of the Neuroendocrine Hypothalamus 11
Molecular studies begin to reveal how chick bHyp cells develop and give rise to
anterior and mammillary basal progenitors. These studies, and studies that suggest
the conservation of anisotropic growth mechanisms in other species, have been
described in detail in a recent review (Fu et al. 2019) and here we provide only a
summary.
12 M. Placzek et al.
Fig. 1.4 Successive stages in bHyp cell induction and anterior progenitor growth and differentia-
tion from bHyp cells. Schematic sagittal views of chick embryos (8–17 somites). (a) Induction of
Shh(+) RDVM cells in the 8 somite embryo. (b) Establishment of dorsoventral pattern through a
Shh morphogen gradient: inset shows patterned progenitor domains. (c) Differentiation of Shh(+)
RDVM cells to bHyp cells that co-express Shh, BMP7 and Fgf10 (red area). (d) Resolution of bHyp
into anterior Fgf10/Shh(+) cells (polka dot) and posterior Fgf10/BMP7(+) (red) domains and onset
of growth (depicted by curved arrow). Shh(+) neuroepithelial progenitors (hatched) that
downregulate Fgf10 grow anteriorly from Fgf10/Shh(+) cells. Note although shown in sagittal
view, Shh(+) neuroepithelial progenitors form in an arc around Fgf10(+) cells. (e) Continued
generation and differentiation of anterior progenitors (orange: from Fu et al. 2017)
bHyp cells develop from a set of Shh(+) cells, termed rostral diencephalic ventral
midline (RDVM) cells that are induced in the ventral midline of the developing
forebrain (prior to the 10-somite stage in chick). Induction of RDVM cells is
mediated by the signalling ligands Shh and Nodal (reviewed in Placzek and Briscoe
2005), deriving from the underlying prechordal mesendoderm (Fig. 1.4a). At their
anterior limit, RDVM cells abut telencephalic progenitor cells (Fig. 1.4a). Tissue
1 Development of the Neuroendocrine Hypothalamus 13
explant studies in chick suggest that Shh diffuses out of RDVM cells and forms a
spatial morphogen gradient in neighbouring diencephalic neuroepithelial cells
(Fig. 1.4b). This results in the establishment of a dorsoventral pattern. Progenitors
that express the homeodomain transcription factor (TF) Nkx2.2 abut Shh(+) RDVM
cells, and those expressing the homeodomain TF Pax6 are established adjacent to the
Nkx2.2(+) domain (Fig. 1.4b inset). The induction of RDVM cells and their role in
establishing an early dorsoventral pattern appear to be evolutionarily conserved.
Genetic lineage-tracing studies in the mouse (Newman et al. 2018b) that build on
ground-breaking large-scale transcriptomic analyses (Shimogori et al. 2010) indicate
that the hypothalamus is of diencephalic origin. Knockout studies, or analysis of
mutated genes, indicate that the prechordal mesendoderm induces Shh(+) RDVM
cells in mouse, zebrafish, and humans, elegant work in mouse showing that Shh
expression in RDVM cells is driven through a unique enhancer, SBE2 (Shh brain
enhancer 2) (Jeong et al. 2006). Genetic studies in mouse and zebrafish show that
Shh that derives from RDVM cells establishes Nkx2.2(+) and Pax6(+) progenitor
domains. If Shh is selectively deleted in RDVM cells (ShhΔhyp mice; Zhao et al.
2012), the Shh-target gene Gli1 is absent, Nkx2.2 is reduced and Pax6 expands
ventrally (Corman et al. 2018). Cross-repressive TF interactions may then sustain
pattern: in zebrafish a ventral expansion of pax6 is detected after elimination/
reduction of nkx2 genes (Manoli and Driever 2014).
As homeodomain progenitor domains become established adjacent to RDVM
cells the prechordal mesendoderm changes its properties, downregulating Shh and
upregulating BMP7 (Fig. 1.4b). Tissue explant studies in chick suggest that BMP7
from the prechordal tissue has a major influence on overlying RDVM cells, causing
them to upregulate BMP7 and then Fgf10, i.e. to become bHyp cells (Fig. 1.4c;
Manning et al. 2006). Upregulation of BMP7/Fgf10 causes the cells to dramatically
change their behaviour, switching from a quiescent to a proliferative state (Manning
et al. 2006; Fu et al. 2017).
Table 1.2 Shh-Rx orchestrate growth and differentiation of the anterior basal hypothalamus
Species Experiment Result References
Zebrafish Rx3 (chokh) Loss of Shh+pea3- progenitors; Tessmar-Raible et al.
mutant or rx3 expansion of basal pea3+ domain; (2007); Dickmeis et al.
morpholino expansion of alar pax6+ domain; loss (2007); Muthu et al.
of anterior and tuberal neurons, (2016)
including pomc, sf1/ff1b, avp, Otp
and TH+ neurons
Mouse Rx2 Loss/reduction of anterior territory Orquera et al. (2016)
conditional (Lhx1, Nkx2.2); loss of anterior and
ablation tuberal neurons, including pomc, Otp,
TH and Sst+ neurons
Chick Timed Reduced Six3+p57+ anterior Fu et al. (2017)
cyclopamine progenitor territory; reduced tract of
(HH9–12) post-optic commissure
Zebrafish Timed Loss of anterior and tuberal neurons Muthu et al. (2016)
cyclopamine including pomc and sf1/ff1b
(30hpf)
Mouse Shh deletion Loss of tuberal and anterior markers Shimogori et al. (2010)
~E10.5 (Lhx1,Lhx6, Lhx9) and neurons
(Nkx2.1-Cre (pomc, sf1)
driver)
Mouse Shh deletion Loss/reduction of anterior markers Zhao et al. (2012);
~9.0 (SBE2- (Six6, Nkx2.1, Nkx2.2); expansion of Corman et al. (2018)
Cre driver) Pax6; loss of tuberal neurons (Hmx3,
Sf1, Pomc, TH, Sst) and reduced
mitotic progenitors
Mouse Shh deletion Loss of anterior territory (Nkx2,1, Carreno et al. (2017)
~E10 (Hesx1- Wnt5a, Tcf4); loss of anterior and
Cre driver) tuberal neurons (GHRH, Pomc, Sst,
Oxt and Avp)
Mouse Shh deletion Reduced Foxd1, Nkx2.1; reduces Szabó et al. (2009)
~E9.0 (Foxb1- anterior/tuberal neurons: Pomc, Hcrt,
Cre driver) reduced mitotic progenitors
Summary of experiments that knockdown, or conditionally knockdown/reduce Rx/rx3 or Shh in
hypothalamic neuroepithelial cells
mesendoderm and acts with Nodal to induce RDVM cells (Fig. 1.4a). Then it derives
from RDVM cells and acts as a morphogen to establish an early dorsoventral pattern
(Fig. 1.4b). Finally (at least in chick and zebrafish: Muthu et al. 2016; Fu et al. 2017),
it derives from Shh(+) neuroepithelial cells to drive growth of the anterior basal
hypothalamus, potentially by regulating cell cycle. Although not yet proved, it
remains possible that Shh(+) neuroepithelial cells continue to act
non-autonomously to pattern adjacent cells. This raises the possibility that progenitor
cells in the basal hypothalamus are continually generated through two routes: from
cells that are responsive to Shh signalling, but do not express Shh; or directly from
Shh(+) neuroepithelial cells, a notion supported through recent work in mouse
(Corman et al. 2018).
16 M. Placzek et al.
Fig. 1.6 (a) Development of the hypothalamo–pituitary axis. The adenohypophyseal (anterior
pituitary) placode (black hatched) moves ventrally (curved arrow) beneath bHyp cells (red). (b) The
placode invaginates to form a pouch as anterior neuroepithelial Shh(+) progenitors (red hatched)
form. (c) Lhx3 is upregulated, forming the definitive RP as anterior progenitors (orange) form. (d) A
subset of bHyp cells give rise to progenitors that grow ventrally and form the infundibulum, the
precursor of the median eminence (ME) and posterior pituitary (PP). RP gives rise to the anterior
pituitary (AP). Inset shows high power view of these regions. Dotted circle in (c, d) denotes optic
stalk
For many days, Fgf10+ive progenitor cells, and their infundibular daughters, are in
intimate contact with the prospective anterior pituitary (AP, aka adenohypophyseal)
cells. The AP derives from a placode that invaginates to form a rudimentary pouch
that differentiates into Rathke’s pouch, defined through the expression of Lhx3
(a master transcriptional regulator of the endocrine cells of the AP). Sequential
22 M. Placzek et al.
Cells within the infundibulum, and at its neck, continue to express Fgfs, including
Fgf10 and Fgf3, long after the major neurogenic periods of hypothalamic develop-
ment and play a role in guiding neurosecretory axon terminals and portal capillary
vessels to set up the glial–neuronal–capillary interfaces that underlie the future
functioning of the hypothalamo–pituitary axis and homeostatic balance (Fig. 1.1).
These studies have been reviewed recently (Burbridge et al. 2016) and here we
provide only a summary.
Genetic studies in mouse first indicated that Fgf signalling promotes the guidance
of parvocellular neurosecretory axons (Tsai et al. 2011). Analyses of fgf3 /
zebrafish then revealed that avp(+) and oxt(+) magnocellular cells, and dopaminer-
gic (DA) parvocellular axons fail to project to the ME-like neurohaemal region or
PP, despite neurons initially developing normally (Gutnick et al. 2011; Liu et al.
2013). Expression of a dominant-negative Fgf receptor, in a temporally- or spatially
controlled manner, revealed that Fgf3 directly affects growing axons. Parallel studies
in chick demonstrated that Fgf3/Fgf10 deriving from the infundibulum act as long-
range chemoattractants: low levels of Fgf3/Fgf10 stimulate and orient the growth of
both AVP(+) and DA neurons. Intriguingly, high concentrations of Fgfs stall or repel
growing axons, a mechanism that might help to ensure that neuroendocrine axons
project to, but do not cross the ventral midline: uniquely in this region of the CNS,
axons are non-commissural. Currently, little is known about subsequent mechanisms
that ensure the correct projection of parvocellular axons to the median eminence, and
1 Development of the Neuroendocrine Hypothalamus 23
of magnocellular axons to the posterior lobe. Future studies are needed to examine
this, and to understand how Fgf ligands operate in conjunction with other factors,
including Notch, Shh, BMP and Netrin (reviewed in Burbridge et al. 2016).
Concomitant with guiding axons, FGFs act with other factors to promote local
endothelial vasculogenesis (Liu et al. 2013). These studies point to a role for FGF
signalling during the early modelling of endothelial cells, and the initial formation of
the capillary plexi that characterize the ME and PP (Gutnick and Levkowitz 2012,
Fig.1.1) and suggest that, by attracting axons and stimulating endothelial cells, Fgfs
ensure the coordinated growth of neuroendocrine axons and capillaries to the same
general target. This is then another step in the self-organizing hypothalamic pro-
gram, one that builds the neurovascular interface over time. Elegant studies in
zebrafish show that as oxt(+) magnocellular neurons innervate the PP, they release
oxt from their endfeet. Oxt acts as an angiogenic cue for nearby vascular sprouts,
drawing them towards axon terminals, where they go on to form the hypophyseal
arteries and veins (Gutnick et al. 2011). Thus, parvocellular axons actively partici-
pate in building the vascular networks required for their activity, another proof-of-
principle for the idea that the hypothalamus self-organizes over time.
In addition to their pivotal roles in acute homeostasis (outside of the scope of this
review: see Table 1.1), Fgf10(+) tanycyte subpopulations have been implicated in
longer-term allostatic change through their ability to act as neural stem/progenitor
(NSP) cells.
EdU-labelling studies in postnatal animals first showed that regions around the
ME are hotspots of neurogenesis and gliogenesis, and subsequent studies revealed
that β1, β2 and α2 tanycyte subsets can act as neurogenic and gliogenic NSP cells
(summarized in Rizotti and Lovell-Badge 2017; Lewis and Ebling 2017; Ebling and
Lewis 2018; Clasadonte and Prevot 2018). Lineage-tracing studies showed that
Fgf10(+) tanycytes, although largely quiescent in adulthood, can respond and
adapt to their changing environment by producing new neurons. Thus, Fgf10(+)
tanycytes respond to a high-fat diet by generating neuropeptide Y (NPY)-neurons
that regulate feeding behaviour (Haan et al. 2013). Other lineage-tracing studies
showed that an FGF-responsive α-2 tanycyte subset can self-renew or give rise to
other tanycyte populations, even in unstimulated mice, indicating that localized
proliferation and differentiation to tanycytes maintains the integrity of the ventricular
wall (Robins et al. 2013). Fgf10 and Fgf signal components are maintained in
subsets of tanycytes (including β- and α-2 subsets) throughout life (Haan et al.
2013; Robins et al. 2013) and Fgf signalling can regulate tanycyte proliferation:
infusion of FGF2 into the third ventricle stimulates proliferation of both β-1 and α-2
subtypes (Robins et al. 2013). Tanycyte neurogenesis and gliogenesis can be
stimulated by a range of signals, hormones and metabolites, and ongoing studies
are now seeking to determine whether/how Fgf signalling is involved in supporting
adult neurogenic programs. One study suggests that tanycyte stimulation by FGF1
correlates with the sustained remission of diabetic hyperglycaemia in rodent models
(Scarlett et al. 2016), suggesting that appropriate regulation of Fgf signalling in
tanycytes may be critical to health and disease. Important questions for the future
are: (1) whether adult programs of tanycyte homeostasis and neurogenesis will share
features of embryonic Fgf10(+) progenitor homeostasis and neurogenesis;
(2) whether physiological or epigenetic factors that are known to govern
neurogenesis from adult tanycytes do so via interactions with developmental
signal-TF pathways; and (3) whether neuroendocrine neurons can be generated
throughout life.
Key References
Carreno et al. (2017)—Genetic studies in the mouse reveal the integrated develop-
ment of the ventral hypothalamus (including the infundibulum and Shh(+)
neuroepithelial progenitors) and the adenohypophysis.
Chen et al. (2017)—A large scale single cell RNA seq analysis of the adult mouse
hypothalamus confirms that neuroendocrine neurohormones/neurotransmitters
are expressed in heterogeneous cell types.
Corman et al. (2018)—Sophisticated mouse studies, including analyses of condi-
tional knock-out and reporter lines, show that tuberal neurons derive from both
Shh-expressing and Shh-responsive progenitors.
Fu et al. (2017)—Fate-mapping and pharmacological studies in the early chick
provides evidence for the anisotropic growth model of hypothalamic
development.
Haan et al. (2013)—Evidence for de novo adult neurogenesis from Fgf10(+)
tanycytes; together with Fu et al. (2017) and Robins et al. (2013), raises possibil-
ity that an Fgf10(+) population is important for building hypothalamic neurons
throughout life.
Kim et al. (2019)—A new study that sets a benchmark for molecular analysis of the
developing hypothalamus. Single cell RNA seq analyses of the hypothalamus at
different time points in embryogenesis and postnatal life provide evidence for the
early specification of key hypothalamic neuronal subsets. Comparison of wild
type and mutant mice provides evidence that the hypothalamus is a diencephalic-
derived structure.
Manning et al. (2006)—A study in the chick embryo that provided an understanding
of the co-ordination of pattern and proliferation of ventral hypothalamic progeni-
tor cells. Re-interpretation of this paper, in light of the anisotropic growth shown
in Fu et al. (2019), suggests that the cells that express Tbx2/BMP are bHyp cells,
and the Emx2+ cells are mammillary progenitors.
Muthu et al. (2016)—Anisotropic growth of anterior hypothalamus in zebrafish;
absence of growth leads to failure of differentiation of both ‘anterior’ and
‘tuberal’ neuroendocrine neurons.
Newman et al. (2018a)—Sophisticated conditional knock-out and reporter lines
show hypothalamus is of diencephalic origin.
Orquera et al. (2016)—Evidence for conserved mechanism of anterior progenitor
growth and differentiation, via a Shh-Rax pathway.
Robins et al. (2013)—Analysis of adult mouse shows that Fgf10(+) stem/progenitor
cells exist as tanycytes that line the ventricle; with Fu et al. (2019) and Haan et al.
(2013), suggests that Fgf10(+) stem/progenitor cells are retained through life.
Shimogori et al. (2010)—A large-scale screen and in situ analysis identifies many
hypothalamic progenitor domains; conditional knock-out of subsets of Shh(+)
neuroepithelial progenitors shows importance in anterior/tuberal growth and
neurogenesis.
Szabó et al. (2009)—Lineage-tracing of a subset of Shh(+) RDVM cells shows
contribution to tuberal and mammillary hypothalamus.
Tessmar-Raible et al. (2007)—Conservation of neuroendocrine cells and conserva-
tion of key molecules.
1 Development of the Neuroendocrine Hypothalamus 27
References
Alvarez-Bolado G (2018) Development of neuroendocrine neurons in the mammalian hypothala-
mus. Cell Tissue Res. https://doi.org/10.1007/s00441-018-2859-1
Alvarez-Bolado G, Paul FA, Blaess S (2012) Sonic hedgehog lineage in the mouse hypothalamus:
from progenitor domains to hypothalamic regions. Neural Dev 7:4. https://doi.org/10.1186/
1749-8104-7-4
Anbalagan S, Gordon L, Blechman J, Matsuoka RL, Rajamannar P, Wircer E, Biran J, Reuveny A,
Leshkowitz D, Stainier DYR, Levkowitz G (2018) Pituicyte cues regulate the development of
permeable neuro-vascular interfaces. Dev Cell. https://doi.org/10.1016/j.devcel.2018.10.017
Bedont JL, Blackshaw S (2015) Constructing the suprachiasmatic nucleus: a watchmaker's per-
spective on the central clockworks. Front Syst Neurosci 9:74
Bedont JL, Newman EA, Blackshaw S (2015) Patterning, specification, and differentiation in the
developing hypothalamus. Wiley Interdiscip Rev Dev Biol. https://doi.org/10.1002/wdev.187
Brinkmeier ML, Potok MA, Davis SW, Camper SA (2007) TCF4 deficiency expands ventral
diencephalon signaling and increases induction of pituitary progenitors. Dev Biol 311:396–407
Burbridge S, Stewart I, Placzek M (2016) Development of the neuroendocrine hypothalamus.
Compr Physiol 6(2):623–643
Campbell JN, Macosko EZ, Fenselau H et al (2017) A molecular census of arcuate hypothalamus
and median eminence cell types. Nat Neurosci 20:484–496
Carreno G, Apps JR, Lodge EJ, Panousopoulos L, Haston S, Gonzalez-Meljem JM, Hahn H,
Andoniadou CL, Martinez-Barbera JP (2017) Hypothalamic sonic hedgehog is required for cell
specification and proliferation of LHX3/LHX4 pituitary embryonic precursors. Development
144(18):3289–3302
Chen R, Wu X, Jiang L, Zhang Y (2017) Single-cell RNA-Seq reveals hypothalamic cell diversity.
Cell Rep 18(13):3227–3241
Clasadonte J, Prevot V (2018) The special relationship: glia-neuron interactions in the neuroendo-
crine hypothalamus. Nat Rev Endocrinol 14(1):25–44. https://doi.org/10.1038/nrendo.2017.124
Corman TS, Bergendahl SE, Epstein DJ (2018) Distinct temporal requirements for Sonic hedgehog
signaling in development of the tuberal hypothalamus. Development 145(21). https://doi.org/
10.1242/dev.167379
Couly GF, Le Douarin NM (1985) Mapping of the early neural primordium in quali-chick chimeras.
Dev Bioll 110:422–439
Davis SW, Ellsworth BS, Peréz Millan MI, Gergics P, Schade V, Foyouzi N, Brinkmeier ML,
Mortensen AH, Camper SA (2013) Pituitary gland development and disease: from stem cell to
hormone production. Curr Top Dev Biol 106:1–47
Dickmeis T, Lahiri K, Nica G, Vallone D, Santoriello C, Neumann CJ, Hammerschmidt M, Foulkes
NS (2007) Glucocorticoids play a key role in circadian cell cycle rhythms. PLoS Biol 5:e78
Ebling FJP, Lewis JE (2018) Tanycytes and hypothalamic control of energy balance. Glia 66
(6):1176–1184
Fu T, Towers M, Placzek MA (2017) Fgf10+ progenitors give rise to the chick hypothalamus by
rostral and caudal growth and differentiation. Development 144(18):3278–3288
28 M. Placzek et al.
homeodomain factor Lhx2 is required for hypothalamic tanycyte specification and differentia-
tion. J Neurosci 34:16809–16820
Samms RJ, Lewis JE, Lory A, Fowler MJ, Cooper S, Warner A, Emmerson P, Adams AC, Luckett
JC, Perkins AC, Wilson D, Barrett P, Tsintzas K, Ebling FJ (2015) Antibody-mediated inhibi-
tion of the FGFR1c isoform induces a catabolic lean state in siberian hamsters. Curr Biol 25
(22):2997–3003
Sánchez-Arrones L, Ferrán JL, Rodríguez-Gallardo L, Puelles L (2009) Incipient forebrain
boundaries traced by differential gene expression and fate mapping in the chick neural plate.
Dev Biol 335:43–65
Scarlett JM, Rojas JM, Matsen ME, Kaiyala KJ, Stefanovski D, Bergman RN, Nguyen HT,
Dorfman MD, Lantier L, Wasserman DH, Mirzadeh Z, Unterman TG, Morton GJ, Schwartz
MW (2016) Central injection of fibroblast growth factor 1 induces sustained remission of
diabetic hyperglycemia in rodents. Nat Med 22(7):800–806
Seth A, Culverwell J, Walkowicz M, Toro S, Rick JM, Neuhauss SC, Varga ZM, Karlstrom RO
(2006) belladonna/(Ihx2) is required for neural patterning and midline axon guidance in the
zebrafish forebrain. Development 133(4):725–735
Shimogori T, Lee DA, Miranda-Angulo A, Yang Y, Wang H, Jiang L, Yoshida AC, Kataoka A,
Mashiko H, Avetisyan M, Qi L, Qian J, Blackshaw S (2010) A genomic atlas of mouse
hypothalamic development. Nat Neurosci 13:767–775
Swanson LW (1992) Brain maps: structure of the rat brain. Elsevier, New York
Szabó NE, Zhao T, Cankaya M, Theil T, Zhou X, Alvarez-Bolado G (2009) Role of neuroepithelial
Sonic hedgehog in hypothalamic patterning. J Neurosci 29(21):6989–7002. https://doi.org/10.
1523/JNEUROSCI.1089-09.2009
Tessmar-Raible K, Raible F, Christodoulou F, Guy K, Rembold M, Hausen H, Arendt D (2007)
Conserved sensory-neurosecretory cell types in annelid and fish forebrain: insights into hypo-
thalamus evolution. Cell 129:1389–1400
Trowe MO, Zhao L, Weiss AC, Christoffels V, Epstein DJ, Kispert A (2013) Inhibition of Sox2-
dependent activation of Shh in the ventral diencephalon by Tbx3 is required for formation of the
neurohypophysis. Development 140:2299–2309
Trudel E, Bourque CW (2010) Central clock excites vasopressin neurons by waking osmosensory
afferents during late sleep. Nat Neurosci 13:467–474
Tsai PS, Brooks LR, Rochester JR, Kavanaugh SI, Chung WC (2011) Fibroblast growth factor
signaling in the developing neuroendocrine hypothalamus. Front Neuroendocrinol 32:95–107
Wang X, Kopinke D, Lin J, McPherson AD, Duncan RN, Otsuna H, Moro E, Hoshijima K,
Grunwald DJ, Argenton F, Chien CB, Murtaugh LC, Dorsky RI (2012) Wnt signaling regulates
postembryonic hypothalamic progenitor differentiation. Dev Cell 23:624–636
Xu C, Fan CM (2007) Allocation of paraventricular and supraoptic neurons requires Sim1 function:
a role for a Sim1 downstream gene PlexinC1. Mol Endocrinol 21(5):1234–1245
Zhang L, Mathers PH, Jamrich M (2000) Function of Rx, but not Pax6, is essential for the formation
of retinal progenitor cells in mice. Genesis 28:135–142
Zhao T, Szabo N, Ma J, Luo L, Zhou X, Alvarez-Bolado G (2008) Genetic mapping of Foxb1-cell
lineage shows migration from caudal diencephalon to telencephalon and lateral hypothalamus.
Eur J Neurosci 28:1941–1955
Zhao Y, Mailloux CM, Hermesz E, Palkóvits M, Westphal H (2010) A role of the LIM-homeobox
gene Lhx2 in the regulation of pituitary development. Dev Biol 337(2):313–323. https://doi.org/
10.1016/j.ydbio.2009.11.002
Zhao L, Zevallos SE, Rizzoti K, Jeong Y, Lovell-Badge R, Epstein DJ (2012) Disruption of SoxB1-
dependent Sonic hedgehog expression in the hypothalamus causes septo-optic dysplasia. Dev
Cell 22:585–596
Sonic hedgehog in Hypothalamus
Development 2
Gonzalo Alvarez-Bolado
Abstract
Sonic hedgehog (SHH) is a secreted signaling protein able to confer ventral cell
fate to the spinal cord on the basis of a concentration gradient that translates into
activation of specific genes through Gli transcription factors. The intricate and
dynamic expression pattern of Shh in the forebrain raises the question of its
function in the developing hypothalamus. Experiments in chick and mouse
show that SHH from the prechordal plate cooperates with Bone Morphogenetic
Proteins (BMPs) to induce a special “hypothalamic floor plate” where Shh
expression will be activated by three transcription factors, Nkx2-1, Six3, and
Sox2/3. Conditional mouse mutants show that this new source of SHH is respon-
sible for hypothalamic growth as well as full differentiation of most hypothalamic
cell types and the formation of the lateral hypothalamus. Experiments on chick
embryos indicate that for the specification of hypothalamic regions Shh
cooperates with Fgf10. For the final differentiation of the basal hypothalamic
regions and hypophysis, repression of Shh by T-box proteins is necessary.
Precisely how SHH unlocks hypothalamic cell fates is not entirely clear. But in
mouse and zebrafish, Shh bestows neuroendocrine cell fates through a complex
relation with Rax/rx3.
Keywords
Floor plate, Gli, Hypothalamus, Lineage, Prechordal plate, Progenitor domain,
Sonic hedgehog
G. Alvarez-Bolado (*)
Department of Neuroanatomy, University of Heidelberg School of Medicine, Heidelberg, Germany
e-mail: [email protected]
2.1 Introduction
Sonic hedgehog (Shh) encodes a secreted protein essential for the development of the
central nervous system (CNS). SHH has fascinated developmental biologists since
its discovery because of its ability to confer a variety of cell fates on the basis of a
concentration gradient. The most important developmental role of SHH in the CNS
is to confer specific character to the ventral half of the neural tube (Ericson et al.
1995). Progenitor cells in the early neural tube (the stem cells of the CNS) receive
their “ventral character” from SHH secreted by a specific source tissue. In the spinal
cord, the first source of SHH is the axial mesoderm underlying the presumptive
spinal cord, i.e., the notochord. Neuroepithelial cells in the ventral midline of the
presumptive spinal cord are induced by SHH to become floor plate cells and start
expressing Shh and secreting SHH themselves into the neural tube. SHH then forms
a concentration gradient effective over a relatively long distance (20 cell diameters
from the source at least) and imparts ventral character to the progenitor cells that
form the early neural tube according to this gradient. The highest concentration
induces the most ventral structures and progressively lower concentrations of SHH
specify progressively less ventral (more dorsal) cell fates. The complex spatial and
temporal properties of the SHH gradient and the ways in which cells respond to it are
well delineated. The resulting body of knowledge is an achievement of modern
developmental biology (Dessaud et al. 2008; Sagner and Briscoe 2017). Most of this
work was done using the spinal cord as a model. In the forebrain, the dynamic
expression pattern of Shh suggested even a more intricate involvement of this
signaling pathway in the development of the ventral forebrain and of the
hypothalamus.
After translation, SHH cleaves itself into an active N-terminal fragment and is then
covalently attached to two lipid moieties, palmitoyl and cholesterol (which probably
confer SHH the peculiar capacity to signal through relatively long distances). SHH is
then released by source cells through one or maybe several mechanisms. Essential
for SHH to exert its influence on the target cell is a cellular compartment, the primary
cilium (Huangfu et al. 2003), as well as several proteins acting as receptors, second
messengers and effectors of the SHH signal. In vertebrates, the SHH transduction
pathway ends on three effector proteins, the GLI transcription factors (GLI1, 2, and
3). Cells competent to respond to SHH express two transmembrane proteins,
Patched (Ptch1) and Smoothened (Smo), as well as Gli2 and Gli3 (and other specific
genes that regulate the pathway).
While GLI1 is a transcriptional activator, GLI2 and GLI3 can act as activators
(GliA) or repressors (GliR). The GLI DNA-binding domain consists of five zinc
fingers, and the three GLI proteins have similar DNA-binding specificity. The
activator domain is the C-terminal, and the repressor domain of GLI2 and GLI3 is
the N-terminal. In SHH responsive cells, GLI2 and GLI3 are translated and
2 Sonic hedgehog in Hypothalamus Development 33
processed into their repressor forms and repress Shh pathway targets (the primary
repressor is GLI3). Upon binding of SHH to the receptor complex, PTCH1 allows
SMO to change into its active conformation. SMO then activates a signaling cascade
that inhibits processing of GLI proteins into their repressor forms and instead they
are processed into their activator forms (the activator form is mostly GLI2). In
contrast, Gli1 is the first gene expressed upon activation of the pathway and therefore
its expression is a readout of Shh pathway activation. (Note: This is somewhat
divergent in fishes). The components of the SHH transduction pathway are localized
to the cilia, possibly allowing for greater sensitivity of recipient cells to SHH
molecules. The SHH signaling cascade is still being unraveled, and the subject has
been often and expertly reviewed (see for instance Briscoe and Thérond 2013).
The adult hypothalamus has been classically subdivided into four transverse regions,
from rostral to caudal: preoptic, anterior, tuberal, and mammillary (Fig. 2.1a).
However, developmental gene expression patterns do not entirely conform to the
neuroanatomy of the adult brain as it is traditionally understood (Puelles et al. 2012)
(Fig. 2.1b; from Haddad-Tóvolli et al. 2015). If we take Shh expression (pink in
Fig. 2.1b) as the ultimate ventral marker, the tuberal and mammillary regions are
ventral (i.e., they express Shh during development) and the preoptic and anterior are
dorsal (Fig. 2.1b). To avoid the resulting “clash of nomenclatures” (i.e., adult versus
developmental neuroanatomy), the words “alar” and “basal” (Fig. 2.1b, yellow
versus red, separated by a red line) are sometimes used instead of “dorsal” and
“ventral” when speaking of the developing hypothalamus. The well-known term
“ventral forebrain” becomes “anterior forebrain” and includes alar (dorsal) and basal
(ventral) portions. The most anterior region has received the name of “acroterminal
region” (green in Fig. 2.1b), the basal portion of which (green striped area in
Fig. 2.1b) corresponds to the Rostral Diencephalic Ventral Midline cells (RDVM;
also called hypothalamic floor plate and Medial Progenitor Domain (used through-
out this chapter)).
For this reason, it is preferable to substitute “presumptive hypothalamus” for
“ventral forebrain.” Even more, developmental studies show that the preoptic region
originates in the telencephalon and is not part of the diencephalon (see Sect. 2.7.3).
These questions reflect the current discussion about the developmental and evolu-
tionary origins of the four hypothalamic regions and are not yet settled.
34 G. Alvarez-Bolado
Fig. 2.1 Hypothalamic regional nomenclature and Shh expression. Conventional representation of
the hypothalamus (gray) as ventral region with four regions aligned from rostral to caudal: POA,
preoptic; ANT, anterior; TUB, tuberal; MAM, mammillary. Model of the hypothalamus consider-
ing Shh expression (pink) as basal (ventral) marker. The tuberal and mammillary regions would
then be not “caudal” but rather basal (“ventral”), and the anterior region would form the alar
(“dorsal”) hypothalamus (yellow). In this model, the most anterior portion of the hypothalamus is
called acroterminal region (green); its basal portion (striped green) corresponds to the Medial
Progenitor Domain or hypothalamic floor plate. Additionally, the preoptic region is considered
part of the telencephalon. The red line represents the boundary of Shh expression. ac, anterior
commissure; hp, hypophysis; MB, midbrain; os, optic stalk; POA, preoptic region; PT, pretectum;
PTh, prethalamus; TH, thalamus; ZLI, zona limitans (modified from Haddad-Tóvolli et al. 2015).
(continued)
2 Sonic hedgehog in Hypothalamus Development 35
At E7.5 (neural fold), Shh is expressed in the prechordal plate, immediately under-
lying the midline of the neural plate (prechordal Shh; (Aoto et al. 2009; Echelard
et al. 1993)) (Fig. 2.2a). This is followed immediately by the onset of Gli1 expres-
sion in the midline of the overlying neural ectoderm, indicating activation of Shh
signaling. Gli1 expression appears in the midline; Gli2 and Gli3, on the contrary, are
expressed widely in the ectoderm, indicating competence to respond to secreted
SHH (Hui et al. 1994).
At E8.5 Shh is expressed in a medial territory overlying the prechordal plate and
extending through the presumptive mammillary and tuberal regions (the two regions
forming the “basal hypothalamus,” Sect. 2.1.2). This territory is the “Medial Pro-
genitor Domain” (Sect. 2.3). This region is characterized not only by Shh expression
like the floor plate in the spinal cord, but for the concomitant expression of Nkx2-1
and Bmp7 genes (while characteristic floor plate marker Foxa2 is expressed here
only transiently) (Dale et al. 1997) (Sect. 2.6.1). Expression of Gli1 and Gli2 overlap
forming a paired lateral band (Fig. 2.2b), while Gli3 is expressed in a peripheral,
non-hypothalamic domain.
At E9.5 Shh expression in the ventral midline extends in the rostral direction but
starts to be downregulated at the tuberal and mammillary level. In this way, a medial
36 G. Alvarez-Bolado
Fig. 2.2 Shh and Gli expression in the developing forebrain. In situ hybridization detection of Shh
(left) and diagrams showing expression domains of the Gli factors and Shh in the presumptive
hypothalamus at the early, middle, and late phases. “lat” and “med,” lateral and medial domains,
respectively. Diagrams (right side) of the corresponding Gli expression patterns with data from
Haddad-Tóvolli et al. (2015) and Hui et al. (1994). The ages indicated (mouse embryonic days, E)
are approximate (since they are based on an estimation of the time of fertilization). Note: Very
similar expression patterns have been found in the mouse, chick, and zebrafish embryos. (a) Shh
expression in the prechordal plate seen on a transverse section of the head folds of an E8.0 embryo.
(b–d) In situ hybridization detection of Shh on wild-type mouse brains. In (c) and (d), the heads
have been sagittally halved to show the neuroepithelium–ventricular zone on the inner side of the
brain; rostral to the left. (b) Shh expression domain in the ventral forebrain at E8.5. (c) Shh
expression is downregulated in the basal hypothalamus at E10.5. (d) Shh expression at E12.5.
Asterisk (), basal hypothalamus; CTX, presumptive cortex; FB, forebrain; HB, hindbrain; MB,
midbrain; os, optic stalk; pm, prechordal mesoderm; PTh, prethalamus; Th, Thalamus; TL, telence-
phalic Shh expression domain (modified from Blaess et al. 2014 and Haddad-Tóvolli et al. 2015)
2 Sonic hedgehog in Hypothalamus Development 37
ventral domain without Shh expression originates, while a lateral band of Shh
expression (the Lateral Progenitor Domain; Sect. 2.3) appears (Fig. 2.2c). This is
the “intriguing” and “dynamic” change in pattern that first attracted attention toward
the role of Shh in the presumptive hypothalamus. The mechanistic explanation for it
was obtained in the chick embryo and revealed one fundamental role of Shh
(Manning et al. 2006) (Sects. 2.6.4.1 and 2.7.2). The lateral band of expression
has been called the “hypothalamic basal plate” (as opposed to the “hypothalamic
floor plate” where Shh was originally expressed). In chick embryos, these cells
migrate “forward” from the diencephalon–mesencephalon boundary (Manning
et al. 2006). A telencephalic domain of Shh expression appears at this stage. This
is the Telencephalic Progenitor Domain and generates neurons for the cortex and
basal ganglia as well as astrocytes for the preoptic region (Sect. 2.3).
At E10.5 (Fig. 2.2c), Gli2 expression is now absent in the presumptive hypothala-
mus. Gli3 and Shh-activation diagnostic marker Gli1 overlap in the medial domain,
indicating an activator function of Gli3 (Gli3A) in the midline at this stage.
At E12.5 the general pattern of Shh in the forebrain does not change; however, the
telencephalic domain becomes a clear separate domain (Fig. 2.2d). Shh expression in
the forebrain peaks at this stage and starts fading, so that after E13.5 it has almost
disappeared. The Lateral Progenitor Domain generates cells also at this stage (Sect.
2.3).
The neuroepithelial cells expressing Shh are in principle stem cells, the cells that
divide to originate neurons and glia for at least part of the hypothalamus. Which
brain cells are born from Shh-expressing neuroepithelium in the ventral forebrain,
i.e., which hypothalamic cells belong to the Shh lineage? (Alvarez-Bolado et al.
2012). In order to trace this cell lineage in mice, a mouse line carrying two different
alleles was generated (Legué and Joyner 2010). The first allele encoded the
recombinase Cre fused to the estrogen receptor (CreER) under the control of
regulatory sequences of Shh (ShhCreER); the corresponding fusion protein remains
in the cytoplasm, not being able to enter the nucleus if not bound to a specific ligand,
the synthetic drug tamoxifen. The second allele, inserted in a permanently active
locus, encodes an easily identifiable and specific gene (e.g., β-galactosidase
a.k.a. LacZ, or enhanced yellow fluorescent protein, EYFP) protected from activa-
tion by a floxed stop cassette. In mouse embryos carrying these two alleles, every
cell producing SHH also produces the protein CreER. Approximately within 6 h of
tamoxifen administration by injection, CreER translocates into the cell nucleus,
where the Cre recombinase remains active for 24 h, enough time to eliminate the
38 G. Alvarez-Bolado
floxed stop cassette so that the cell (and its progeny) will permanently express
β-galactosidase (or EYFP). To label developmental events in the mouse, tamoxifen
is administered as an intraperitoneal injection to pregnant females at the desired
embryonic day; injection at E8.5 (TM8.5) guarantees labeling of any cell expressing
Shh between E9.0 and E10.0 and its progeny (lineage labeling).
Beyond E9.5, the Medial Progenitor Domain does not express Shh anymore, while
two novel expression domains appear (Fig. 2.2c). One of them, the Lateral Progenitor
Domain, forms a paired lateral band spanning the tuberal and mammillary regions;
these bands meet rostrally and ventrally immediately behind the optic fissure
(Fig. 2.2c). This Shh expression domain has been called the hypothalamic basal
plate. Analysis of mouse mutants shows that the formation of this domain is depen-
dent on Shh expressed by the neuroepithelium (Sect. 2.5; Fig. 2.5g–i). From TM9.5 to
TM12.5, the fate-mapped cells are found in the Dbx1 expression domain in the
tuberal hypothalamus, with the dorsal-most fate-mapped cells in the hypothalamus
2 Sonic hedgehog in Hypothalamus Development 39
Fig. 2.3 Lineage of Shh in the mouse hypothalamus. (a) Sagittal section through the E18.5 mouse
hypothalamus labeled for Shh lineage (X-gal reaction, blue) after tamoxifen injection at E8.5 to
identify cells derived from the Medial Progenitor Domain. (b) Summary of A (Medial Progenitor
Domain) in the adult mouse hypothalamus. Insets on the right side: Transverse sections of E12.5
mouse brains showing the distribution of the fate mapped cells at four different rostrocaudal levels.
(c) Right side of hemi-dissected adult brain labeled for Shh-lineage (X-gal reaction, blue), after
tamoxifen injection at E11.5 (Lateral Progenitor Domain). (d) Summary of C (Lateral Progenitor
Domain) in the adult mouse hypothalamus. Insets on the right side: transverse sections of E12.5
40 G. Alvarez-Bolado
abut the Sim1-positive domain localized to the anterior region. In addition, the
Telencephalic Progenitor Domain starts to become obvious rostral to the eye primor-
dium (Sects. 2.3.4 and 2.7.4). At this stage, the predominant source of Shh in the
presumptive hypothalamus is the forebrain neuroepithelium (as opposed to the
prechordal plate, where Shh expression dwindles after E10.0); this agrees with the
loss of the lateral hypothalamus in neural Shh mutants (Szabó et al. 2009) (Sect. 2.5).
GLI2 is probably the predominant Gli activator in this domain, but in Gli2-deficient
mice GLI3 can partially substitute for it (Sect. 2.4). Finally, beyond E12.0 and up to
E14.0 at least (Sect. 2.2), numerous cells are born for the lateral hypothalamus of the
tuberal region and for the ventromedial nucleus (Fig. 2.3f). Although the general
pattern at this stage is similar as in the previous, the total number of cells generated is
smaller, and most of them are astrocytes. In the adult brain, cells originated in the
Lateral Progenitor Domain are found mostly in the tuberal region at all mediolateral
levels except the ventral midline (Fig. 2.3c, d, f, g). The mammillary, supra-
mammillary, and premammillary dorsal and ventral nuclei originate mostly in the
Medial Progenitor Domain (although part of the posterior hypothalamus derives from
the Lateral Progenitor Domain). As is the case with the Medial Progenitor Domain,
the Lateral produces only a few, scattered cells for the anterior region.
TM12.0–TM14.0 embryos show Shh-derived cells in the preoptic region, which are
mostly astrocytes (identified on the basis of morphology and GFAP expression).
Only very few preoptic neurons are derived from Shh-expressing progenitors; the
large majority of the neurons of the hypothalamic preoptic region does not belong to
the Shh lineage (i.e., are not derived from the Telencephalic Domain of Shh
expression in the mouse) (see Sect. 2.7.4 for a discussion).
Fig. 2.3 (continued) mouse brains showing the distribution of the fate-mapped cells at four
different rostro-caudal levels. The lineage of the telencephalic Shh domain in the hypothalamus
(mostly astrocytes) is labeled in red. (e, f) Diagrams of a transverse section through the adult mouse
tuberal region showing the distribution of fate-mapped cells from the Medial (e) and Lateral (f)
Progenitor Domains. (g) Merged diagrams from (e) and (f) showing clear differential distribution of
early (blue) and late (yellow) fate-mapped cells on the ventral side; at dorsal levels, the respective
patterns of distribution are less clearly demarcated. ac, anterior commissure; ANT, anterior region;
ARH, arcuate nucleus; cc, corpus callosum; DMH, dorsomedial nucleus; LH, lateral hypothalamus;
LM, lateral mammillary nucleus; MAM, mammillary region; ME, median eminence; MM, medial
mammillary nucleus; mth, mammillothalamic tract; ox, optic chiasm; PMD, dorsal premammillary
nucleus; PMV, ventral premammillary nucleus; POA, preoptic region; PVH, paraventricular
nucleus; RCH, retrochiasmatic area; SuM, supramammillary nucleus; VMH, ventromedial nucleus.
In b, d–g, closed circles indicate neurons, open circles indicate astrocytes, and stars indicate
tanycytes (modified from Alvarez-Bolado et al. 2012)
2 Sonic hedgehog in Hypothalamus Development 41
2.3.5 Conclusions
The secreted protein SHH exerts its influence through transcription factors GLI1,
2, and 3, of which GLI2 and GLI3 can act as activators (GliA) or repressors (GliR).
By comparing the hypothalamic phenotypes of mutant mice deficient in Gli2
(Gli2zfd/+, Gli2 Zinc Finger-Deleted; (Mo et al. 1997)), Gli3 (Gli3Xt-J/+, Extra-
Toes; (Hui and Joyner 1993)), or neural Shh and Gli3 (Foxb1-Cre;ShhloxP/+;
Gli3Xt-J/+; see Sect. 2.5), it is possible to deduce which factors act as GliA or GliR
in the Medial and Lateral Progenitor Domains and if there are differences between
them (Haddad-Tóvolli et al. 2015). We have tried to summarize a complex set of
mutant phenotypes in the diagrams in Fig. 2.4.
2.4.1 The Only Role of Shh in the Alar Regions Is to Suppress Gli3R
The preoptic and anterior regions (alar hypothalamus) show strong expression of
Gli3; despite this fact, the overall specification of these regions is not altered in the
Gli3 mouse mutant brain. It could be that, in the alar portions of the hypothalamus,
Gli2A and Gli3A can substitute for each other. Alternatively, in these regions, the
only role of Shh would be to suppress Gli3R; this would agree with observations that
in double Shh/Gli3 mutant mice the dorsoventral patterning of the telencephalon is
appropriate (Rallu et al. 2002), and supports the hypothesis that both the preoptic and
the anterior regions are telencephalic in origin (Sect. 2.7.4.2). Neither Gli2 nor Gli3
is required for the general patterning of the alar (anterior and preoptic) regions of the
hypothalamus.
In the tuberal and mammillary regions (the basal hypothalamus), Gli2 is required for
the development of the Medial Progenitor Domain (Fig. 2.4a) and its derivatives:
median eminence, pituitary, most of the arcuate nucleus and ventromedial nucleus,
and the mammillary and premammillary nuclei. The results show that the Medial
Progenitor Domain requires nonneural Shh acting through Gli2A (this is in agree-
ment with the phenotype of mice deficient in neural Shh (Shimogori et al. 2010;
Szabó et al. 2009)) (Sect. 2.5). In contrast, the Lateral Progenitor Domain does not
require Gli2 (Fig. 2.4a): the lateral part of the ventromedial and the arcuate nuclei as
well as the dorsomedial nucleus and the lateral hypothalamus. This could mean that
GliA is not required by the Lateral Progenitor Domain. More likely, however, is that
Gli3A is able to compensate in the Gli2zfd/zfd mutant. Gli1 expression (diagnostic of
Shh pathway activation) and Gli3 expression overlap in the Lateral Progenitor
Domain of the Gli2zfd/zfd mutants. In mutants deficient in neurally expressed Shh,
Lateral Progenitor Domain derivatives disappear (Fig. 2.4a) indicating that Shh is
2 Sonic hedgehog in Hypothalamus Development 43
Fig. 2.4 Differential contribution of Gli factors to the Lateral and Medial Progenitor Domains. (a)
Summary diagrams of progenitor domains (neuroepithelium) of the basal hypothalamus in wild
type and mutants as deduced from the phenotypic analysis in the present study. (b) Contribution of
Gli proteins to the specification of the medial and lateral progenitor domains in three successive
stages of development. Colored dots in (a) and (b) represent different kinds of Shh-expressing
progenitors, as indicated on the figure; dotted squares in (b) indicate that, although nonneural Shh
determines the lateral domain through repression of Gli3, a direct influence of Shh through Gli2A
on lateral progenitors before E8.5 is in principle possible (since there are some lateral progenitors in
the double mutant deficient in neural Shh and Gli3). The asterisk () means that loss of GliA2 could
be compensated by Gli3A. MBO, mammillary body (modified from Haddad-Tóvolli et al. 2015)
Fig. 2.5 Phenotype of mouse mutants deficient in neural Shh. (a) In situ detection of Shh
expression with a probe detecting exon2, essential for SHH activity; E12.5 wild-type mouse
embryo. (b) Diagram of Shh expression (red); E12.5 Shh cKO-Nkx2-1 mouse embryo. (c) In situ
detection of Shh-exon2; E12.5 Shh cKO-Foxb1 embryo. (d) Diagram of Shh expression (red);
sagittal section of an E12.5 wild type mouse embryo, showing forebrain regions of interest. (e)
Diagram of Shh expression (red); sagittal section of an E12.5 Shh cKO-Nkx2-1 mouse embryo,
showing forebrain regions of interest. (f) Diagram of Shh-exon2 expression (red) on the right half of
the brain of an E12.5 Shh cKO-Foxb1 mouse embryo. (g) In situ detection of Shh-exon2 expression;
E10.5 wild-type mouse embryo. (h) In situ detection of Shh, either intact or truncated (long probe;
see text for details) expression; E10.5 5 Shh cKO-Foxb1 mouse embryo. (i) In situ detection of Shh-
exon2 expression; E10.5 Shh cKO-Foxb1 mouse embryo. (j, k) In situ detection of Lhx1
2 Sonic hedgehog in Hypothalamus Development 45
Gli3 is not required for the overall specification of the hypothalamus in wild-type
mice (although Gli3A is specifically required for mammillary proliferation). How-
ever, the ventral forebrain phenotypes of Shh mutants, be it the full Shh null mutant
(Chiang et al. 1996) or the conditional Shh mutants lacking Shh expression in the
presumptive hypothalamus (“neural Shh mutants”) (Shimogori et al. 2010; Szabó
et al. 2009) (Sect. 2.5), are much more severe than the Gli2zfd/zfd mutant phenotype.
This indicates that suppression of Gli3R function by Shh (of neural or prechordal
origin) is required for normal hypothalamic development. In support of this conclu-
sion, the defects in the Medial and Lateral Progenitor Domains found in neural Shh
mutants are largely rescued in double neural Shh –/–; Gli3 –/– mice (Fig. 2.4a).
Box 2.2: From the Fruit Fly Larva to the Mouse Brain: The Pathway
Leading to the Role of Shh in Hypothalamic Specification
The first hedgehog gene (hh) was described in the fruit fly as a gene involved
in larval segmentation (Nüsslein-Volhard and Wieschaus 1980). Three
homologs of hh were later found in chick, zebrafish, and mouse (Echelard
(continued)
ä
Fig. 2.5 (continued) expression; sagittal sections of E18.5 wild type or Shh cKO-Foxb1 mouse
brains. Arrow in b, c, e, f shows remnant Shh expression domains in the conditional mutants.
Arrows in g, h, and i show the lateral Shh expression domain or its theoretical location (black arrow)
and the telencephalic Shh expression domain (white arrow). Asterisk in e, f, and k marks the
abnormal size of the third ventricle in the mutants. ANT, anterior region; CTX, cortex; EmTh,
thalamic eminence; MAM, mammillary region; MBO, mammillary body; PTh, prethalamus; SCH,
suprachiasmatic nucleus; Th, thalamus; TUB, tuberal region (b, d, e with data from Shimogori et al.
2010)
46 G. Alvarez-Bolado
The first Shh null mutant mouse (Chiang et al. 1996) confirmed the ventralizing role
of Shh in the hypothalamus and extended it to the forebrain. However, the phenotype
of this mutant was dominated by extreme holoprosencephaly, so that the entire
hypothalamus together with most of the forebrain were missing. Obviously, Shh
expression in the prechordal plate (“prechordal Shh”) was indispensable for fore-
brain development. Conditional mutants eliminating Shh expression from the neural
tube (“neural Shh”) but not the prechordal plate have been much more informative.
By “subtracting” the neural Shh phenotype we could also infer the role of prechordal
Shh. The first such neural mutant removed Shh expression in the forebrain by using
Foxb1 regulatory sequences as the Cre-driver (Foxb1-Cre Shh fl/fl or Shh
cKO-Foxb1 (Szabó et al. 2009). The method was used again by using a Nkx2-1 as
Cre-driver (Nkx2-1-Cre ShhloxP/loxP or Shh cKO-Nkx2-1 (Shimogori et al. 2010).
Further observations on the role of Shh in the ventral forebrain were made using a
Cre line driven by SBE2, an upstream regulatory element of Shh specific for the
hypothalamus (Zhao et al. 2012). Finally, mutants of transcription factor Six3 have
thrown light on the complex role of neural Shh in the splitting of the eye field and
telencephalic vesicle (Geng et al. 2008, 2016). These do not pertain directly to
hypothalamic development (except inasmuch as the retina can be considered a
hypothalamic derivative) but will be included here for the sake of completeness.
2 Sonic hedgehog in Hypothalamus Development 47
1. Not necessary to confer basic hypothalamic identity. Prechordal Shh must pro-
vide general hypothalamic specification, although is not enough for full growth
and differentiation
2. Not necessary to specify the four hypothalamic regions and their positions
relative to each other
3. Essential for the hypothalamic regions to differentiate and express their charac-
teristic marker genes
4. Essential for the hypothalamus to reach appropriate size
5. Essential for the establishment of the Lateral Progenitor Domain and therefore for
the development of the lateral hypothalamus (Sect. 2.3). Prechordal Shh can
partially substitute for the Medial Progenitor Domain
6. Not essential for the specification of the neurohypophysis (but required for its
correct development)
7. Essential for the development of the optic disk
8. Not essential for the splitting of the eye field and telencephalic vesicle, although it
has a minor role in this process
both mutants agree, and some differences (but no contradictions) found are probably
due to the Cre-driver genes used having differential expression and leaving different
Shh expression domains intact (either the telencephalic domain or the midbrain
domain), which in turn could partially rescue the wild-type phenotype. In addition,
some of the conclusions reached may be reflective of the different ages that were
analyzed.
Analyzing mutant mouse lines carrying mutations that decrease Shh in the
neuroepithelium directly or indirectly, have noticed an important role of Shh in
hypophysis development. Zhao et al. (2012) used an SBE2-Cre line to abolish Shh
expression in the presumptive hypothalamus and observed an ectopic, dysmorphic,
and small infundibulum (primordium of the posterior hypophysis) as well as a
duplicated and not completely differentiated Rathke’s pouch (primordium of the
anterior hypophysis). Failure to form the infundibulum as well as a multiplicity of
Rathke’s pouches have been described in mice mutant for Rax, a transcription factor
activated by Shh and, in turn, essential to maintain Shh expression in the presumptive
hypothalamus (Orquera et al. 2016). In mice carrying null mutations of transcription
factor Tbx3, Shh expression abnormally persists in the tuberal region beyond E9.5.
50 G. Alvarez-Bolado
In these embryos, the infundibulum does not develop and Rathke’s pouch shows
decreased proliferation (Trowe et al. 2013). In conditional mutants of Shh using
Hex1 as a Cre driver, Shh expression in the anterior hypothalamus is lost. These
mutants do not develop an infundibulum or Rathke’s pouch (Carreno et al. 2017).
SBE2-Cre ShhloxP/loxP mice lack Shh in the presumptive hypothalamus and show
lack of formation of the optic disk (the point of exit of the retinal axons from the eye)
and a hypoplastic optic nerve (Zhao et al. 2012). Early in the development of the
mammalian head, the prechordal plate is essential for splitting of the eye field and the
primary (unpaired) telencephalic vesicle into left and right. Failure of this process
leads to a major phenotype, holoprosencephaly (cyclopia). Some mouse mutants in
transcription factor Six3, expressed in the rostral portion of the ventral telencephalon,
show holoprosencephaly due to deficiency in Shh in that expression domain (Geng
et al. 2008, 2016). However, none of the three neural Shh mutants (Shimogori et al.
2010; Szabó et al. 2009; Zhao et al. 2012) shows holoprosencephaly.
Experiments on chick embryos have shown that during gastrulation, the prechordal
plate migrates rostrally after detaching from the node. At the same time, the
presumptive hypothalamic floor plate cells migrate rostrally, at times in register
with the prechordal plate, and are influenced by SHH and BMP ligands secreted by
it. If SHH confers “floor plate” identity, the BMPs bestow the hypothalamic floor
plate with “rostral” identity, so that they become “hypothalamic floor plate,” char-
acteristically expressing not only Shh but Nkx2-1 and Bmp7 (Dale et al. 1997; Patten
et al. 2003; Placzek and Briscoe 2005) (summarized in Fig. 2.6a). In experimental
conditions, SHH can by itself rostralize the forebrain floor plate (without the help of
BMP), but only at very high, nonphysiological concentrations (Dale et al. 1997).
Shh secreted by the prechordal plate acts on the overlying neural tube already
patterned along the anteroposterior axis. This early patterning depends on a balance
between the posteriorizing influence of Wnt ligands and the anteriorizing influence
of Wnt antagonists (like Fgf8 from the anterior neural ridge (Shimamura and
Rubenstein 1997)). As a consequence, the forebrain is divided into an anterior
domain expressing transcription factor Six3 and a posterior one expressing Irx3,
meeting each other at the level of the zona limitans (Braun et al. 2003; Kobayashi
et al. 2002). Repression of Wnt1 by Six3 is required to maintain the forebrain
territory and therefore the hypothalamus (Lagutin et al. 2003). Wnt antagonism is
also essential for the ventral forebrain to acquire hypothalamic character (Kapsimali
et al. 2004).The caudalizing Wnt signals (Braun et al. 2003) have to be consistently
2 Sonic hedgehog in Hypothalamus Development 51
Fig. 2.6 Diagrams depicting Shh cooperating and interacting with other genes in the developing
hypothalamus. (a) In most of the neural tube, SHH (orange-colored arrows) from the notochord
(dashed orange-colored line) ventralizes the floor plate (solid orange-colored line), which then
continues secreting SHH; BMP ligands (blue line and arrows) act as dorsalizers. Rostral to the
midbrain–forebrain boundary (dashed black line), the prechordal plate (red dashed line) secretes
SHH plus BMP7, which cooperate to induce the hypothalamic floor plate (red line) (data from Dale
et al. 1997). (b) Shh expression by the hypothalamic floor plate (red) requires Nkx2-1 expression
(pink; at least indirectly, dashed pink arrow) as well as the direct influence (continuous arrows) of
SIX3 (blue) and SOX2/3 (green) non-Shh regulatory sequences. SHH, in turn, maintains Six3 and
Nkx2-1 expression (dashed red arrows) (data from Geng et al. 2008; Sussel et al. 1999; Zhao et al.
2012). (c) The striking change in Shh expression pattern (red) between the middle stage (top left)
and the late stage (bottom right) depends on early secretion of BMP ligands from the prechordal
plate (dashed red line), which activate expression of Bmp genes in the presumptive hypothalamus,
leading to expression of Tbx transcription factors which directly repress Shh expression in a
restricted area (asterisk) (data from Manning et al. 2006; Trowe et al. 2013). (d) Top: The Shh
(red) and Rax (blue) expression domains in the midline of the presumptive hypothalamus by E11.5
overlap in a domain (brown) of the retrochiasmatic area (RCH or ABasM). Bottom: In the zebrafish,
Rax homolog rx3 is activated by Shh (red arrows) in order to confer specific neuroendocrine cell
fates to progenitors; then rx3 is repressed by Shh influence (red lines). The interaction is partially
52 G. Alvarez-Bolado
Fig. 2.6 (continued) conserved in the mouse (black arrows), where Shh is required to activate Rax
expression to accomplish a similar role, and Rax, in turn, maintains Shh expression at a very specific
point in time; repression of Rax by Shh is also conserved in the mouse (data from Lu et al. 2013;
Muthu et al. 2016; Orquera et al. 2016)
2 Sonic hedgehog in Hypothalamus Development 53
cessation of cell proliferation and prevents the specific marker gene Emx2 to be
expressed in the presumptive mammillary region. In this model, it is not possible to
examine subsequent alterations in hypophysis development. In the mouse, Tbx2 is
not required for hypothalamic development, but Tbx3 shows the same interaction
with Shh as in chicken. Tbx2 and Tbx3 could act redundantly, but Tbx2 expression is
lost in the Tbx3 mutant. In the mouse, however, although Tbx3 is essential to repress
Shh expression in the presumptive tubero-mammillary region, the knockout of Tbx3
causes no obvious changes in differentiation (expression of Emx2 and Fgf10 are
normal) but prevents the formation of the infundibulum (presumptive neurohypoph-
ysis, Trowe et al., 2013). In addition, these studies showed that Tbx2 and 3 can
repress Shh expression in this region directly, through two mechanisms. First, both
TBX2 and TBX3 can bind to SOX2 (or SOX3) (necessary to activate Shh), in this
way making it unavailable to bind to the required Shh long-range enhancer SBE2.
Second, both can also specifically bind to long-range enhancer of Shh SFPE2
repressing Shh expression in this region again.
hypothalamic development could be due to lack of neural Shh (since Rax is needed at
E8.5 to keep Shh activated from then on): reduced proliferation leading to thinning
of the midline forebrain neuroepithelium and/or reduced expression of proneural
genes Ascl1 and Ngn3. Additionally, upon losing neural Shh expression (in the Rax
mutant), the rostral expression boundary of markers like Otx2, Tbx3, and Fgf10
shifts rostrally. Perhaps as a consequence, ectopic Rathke’s pouches (presumptive
anterior hypophysis) are found in these mutants (same as in neural Shh mutants
(Zhao et al. 2012)) and the infundibulum is missing (see Sect. 2.7.2).
About the role of Rax in progenitor specification there is some disagreement. One
report including a very detailed description of the Rax expression domain finds that
this gene is not expressed in the presumptive neuroepithelium of the anterior region
(Lu et al. 2013). As a consequence, Rax does not specify neurons of the anterior
region (i.e., Avp- or Oxt-expressing neurons, for instance). By other reports, Rax is
indeed expressed in the presumptive neuroepithelium of the anterior region (Orquera
et al. 2016; Shimogori et al. 2010). Accordingly, Orquera et al. (2016) find a
reduction in the expression of Six6, Lhx1, Nkx2-2, which are markers of the
suprachiasmatic nucleus. Both analyses of Rax mutants find that Rax is essential
for the specification of some progenitors of the arcuate nucleus. Orquera et al. (2016)
report a reduction in markers found in this nucleus of the tuberal region including
Isl1, Pomc, and Th in neuroendocrine neurons as well as Sst (a marker of some
non-neuroendocrine arcuate neurons; reviewed in (Alvarez-Bolado 2018)).
According to Lu et al. (2013), Rax is essential to confer fate specification to the
progenitors of Pomc-expressing neurons in the arcuate nucleus and of SF1 (Nr5a1)-
expressing neurons in the ventromedial nucleus, both in the tuberal region.
Contrary to what happens in zebrafish, in the mouse, Rax is not involved in the
development of the anterior region (i.e., not involved in the specification of Avp-
expressing neurons of the paraventricular nucleus). This may be due to Rax having
an ancestral developmental mechanism in fishes and a phylogenetically more recent
and more restricted function in mammals. This is consistent with the fact that rx3
acts through activation of optb (Otp homolog), a transcription factor essential for
anterior/tuberal region in the fish. In the mouse, however, Otp is specific of anterior
fates, particularly in the paraventricular and supraoptic nuclei (although it also
regulates the development of Sst-expressing neurons of the arcuate nucleus
(Acampora et al. 1999)). For the sake of completeness, it is worth mentioning that
Rax has an important role in tanycyte differentiation (Miranda-Angulo et al. 2014;
Salvatierra et al. 2014) and a tanycyte phenotype has not been reported in Shh
mutants.
2.6.4.4 Fgf10
Both Fgf10 and Shh are expressed in the presumptive hypothalamic neuroepithelium
and the territories of expression of the two markers tend to be complementary (rather
than overlapping), so that they seem to repress each other, in a direct or indirect way.
This expression pattern is key to the development of the hypophysis in chick
embryos (Fu et al. 2017) (see Sect. 2.7.2), and, at least in this last model, seems
key to the specification and growth of the anterior region (see Sect. 2.7.3).
Together with the mammillary, the tuberal region forms the basal portion of the
hypothalamus (Sect. 2.1.2) and most of it derives from Shh-expressing
neuroepithelium (Sect. 2.3). The ventromedial, dorsomedial, and arcuate nuclei are
anatomical landmarks of this region. The arcuate nucleus in particular is easy to
recognize by its functionally important neuronal subpopulations expressing Pomc
and Carptp, Npy, and Agrp or Th and Ghrh. The medial portions of the arcuate and
ventromedial nuclei derive from the medial progenitor domain, and the rest derives
from the Lateral Progenitor Domain (Sect. 2.3). In neural Shh mutants, the tuberal
region is reduced in size and mostly undifferentiated (Sect. 2.5). Transcription
factors essential for tuberal development and depending on neural Shh for activation
or maintenance of expression are Nkx2-1 and Rax (Sects. 2.6.3.1 and 2.6.4.2). Mice
missing Nkx2-1 lack differentiation of the tuberal and the mammillary regions,
which appear to be, however, not much smaller than in wild-type brains (Kimura
et al.1996). Of note, the mammillary region of the Nkx2-1 mutant is more affected
than that of neural Shh mutants, while the anterior region is not affected in Nkx2-1
mutants but very much in neural Shh mutants (see below). In Rax mutant mice, the
arcuate nucleus particularly, and also the ventromedial nucleus, are missing specific
cell populations (Lu et al. 2013; Orquera et al. 2016). The Rax homolog rx3 is
essential for the development of the anterior/tuberal region in zebrafish (Muthu et al.
2016).
of anterior ventral midline cells called “collar cells.” These are characterized by Fgf3
and Sox3 expression and have a key role in the formation of the infundibulum
(Pearson et al. 2011). Next, Shh is repressed by Tbx2 (Tbx3 in mammals) (Manning
et al. 2006; Trowe et al. 2013) in order for the differentiation of the tuberal and
mammillary regions to proceed (Manning et al. 2006), as well as for infundibulum
morphogenesis (Trowe et al. 2013). Therefore, without Shh, no infundibulum
develops. But, if Shh is expressed in the presumptive tuberal–mammillary region
for too long, then no infundibulum develops either. Since Shh expression is
maintained by several factors, in the absence of these, no Shh, therefore no infun-
dibulum: e.g., without Sox3, no Shh, no infundibulum (Zhao et al. 2012); without
Rax, no Shh, no infundibulum (Orquera et al. 2016). Shh is necessary but not
sufficient for hypophysis development, as other factors are also needed. For instance,
as has been observed, in the Nkx2-1 mutant, Shh is expanded, but this does not rescue
the hypophysis phenotype in this mutant (Carreno et al. 2017). Finally, Shh is
required to specify Rathke’s pouch progenitors in the chick (Fu et al. 2017) as
well as in the mouse (Carreno et al. 2017), identified by expression of Lhx3 and
Lhx4. Again, Shh is required but not sufficient to specify them, since BMP and FGF
signaling are also required (Carreno et al. 2017).
The anterior region is in the alar portion of the hypothalamus (Sect. 2.1.2) and its
characteristic nuclei are the paraventricular, the supraoptic, and the suprachiasmatic.
According to mouse mutant phenotypes (Sect. 2.5), the development of the anterior
region is strongly dependent on neural Shh (Sect. 2.3). In zebrafish, transcription
factor rx3 patterns in this region (Muthu et al. 2016). In the mouse, Rax (rx3
homolog) is expressed in the anterior region (Shimogori et al. 2010; Pak et al.
2014) and it could have a role in anterior patterning (Sect. 2.6.4.2). The anterior
region does not derive from Shh-expressing neuroepithelium in the mouse, i.e., it
does not originate in the basal hypothalamus (Alvarez-Bolado et al. 2012). In the
chicken, however, anterior progenitors derive from Fgf10- and Shh-expressing
midline progenitors present in the presumptive tuberal region (Fu et al. 2017). The
most rostral of these progenitors downregulate Fgf10 and move (or proliferate)
forward (or in the alar direction) abandoning the presumptive tuberal region,
downregulate Shh and proliferate intensively, generating the anterior region
(Fu et al. 2017). In this way, the anterior region of the chick would indeed belong
to the Shh lineage.
In both chicken and mouse, the rostral border (alar border) of the Shh-expressing
domain in the presumptive hypothalamus abuts the Foxg1 (BF1) domain that marks
the telencephalon (Fu et al. 2017; Shimamura et al. 1995). Since in the mouse the
anterior region is not Shh-derived, it could be Foxg1-derived, i.e., telencephalic, like
the preoptic region. There is scarce information on the lineage of Foxg1 in the
forebrain, but at least part of the anterior region seems to belong to the Foxg1
lineage, including part of the paraventricular and/or supraoptic nuclei (Kawaguchi
et al. 2016). The Foxg1 mutant mouse shows a phenotype in the magnocellular
58 G. Alvarez-Bolado
neuroendocrine system (Frullanti et al. 2016). Maybe a clue can be found in the fact
that the radial glial fibers that presumably guide the migration of newborn neurons
originate in telencephalon as well as diencephalon and they end in the preoptic and
anterior regions (Tobet et al. 1995).
In terms of neuronal nuclei and connections, the preoptic region is the most complex
and diverse. Furthermore, several of these nuclei are sexually dimorphic. The
preoptic region is involved in the regulation of neuroendocrine secretion (through
numerous efferents to neuroendocrine nuclei of the anterior and tuberal regions),
cardiovascular responses, fluid homeostasis, sleep, body temperature, and sexual
behavior (Simerly 2004). Regulation of preoptic region patterning and development
is not well understood. The morphological complexity and the lack of neuroendo-
crine nuclei with obvious neuropeptide markers make it difficult to study. These
problems are compounded by a dilemma. The preoptic region was considered a part
of the hypothalamus since it was first defined (Le Gros Clark 1938) and certainly, the
hypothalamus is a singular brain region with specific functions (Thompson and
Swanson 2003). However, the preoptic region is now considered part of the telen-
cephalon (Puelles et al. 2012). The problem of the “clash of nomenclatures” between
development and adult has been commented before (Zhang and Alvarez-Bolado
2016). Unfortunately, when dealing with the hypothalamus, the preoptic region is
considered telencephalic and therefore often left out of the scope of the paper
(Shimogori et al. 2010; Fu et al. 2017), while papers dealing with telencephalic
development do not mention the preoptic region since it is hypothalamic (Xu et al.
2008). The net result being that we do not know much about the patterning of this
functionally important region.
The mammillary region is the most caudal (or the most basal) of the hypothalamic
regions. Although it does not have the neuropeptide-expressing neuroendocrine
nuclei that characterizes the anterior and tuberal regions, the mammillary region is
morphologically very distinctive with the large mammillary body and its prominent
axonal tree forming the mammillothalamic and mammillotegmental branches. In the
mouse, the mammillary region starts receiving Shh-expressing medial progenitor
domain cells at E8.0, peaks from E9.0 to E10.0 and reduced numbers of cells keep
being added through E12.0 (Alvarez-Bolado et al. 2012). In the chicken, this process
is different. Fgf10-expressing progenitors in the tuberal region proliferate beyond
their caudal (or basal) boundary; then they downregulate Fgf10, upregulate specific
60 G. Alvarez-Bolado
mammillary marker Emx2, and resume proliferation to form the mammillary region
(a similar process takes place in the anterior region; Sect. 2.7.3) (Fu et al. 2017).
The mammillary body and premammillary nuclei show differential needs for
neural Shh. In the absence of neural Shh in this region, expression of a specific
marker of the mammillary body, Foxb1, is not maintained. Yet, other specific
markers of this nucleus (Emx2, Lhx1, Nkx2-1, Sim1) are still expressed (although
in very reduced domains). The premammillary area, however, is more affected: it
loses specific marker expression at E12.5 and by E18.5 the entire area seems to have
disappeared. Note: in the Nkx2-1 mutant brain, the mammillary region does not
differentiate at all (Kimura et al. 1996), that is, it is much more affected than in neural
Shh mutants. In chick embryos, repression of Shh by Tbx2 in the tubero-mammillary
region is necessary for Emx2 to be expressed, that is, for the differentiation and
further proliferation of specific mammillary progenitors (Manning et al. 2006).
Analysis of Tbx3 mutants shows that, in the mouse, inhibition of Shh expression
in this region is not necessary for Emx2 to be expressed (Trowe et al. 2013). Gli3A is
specifically required for the proliferation of the mammillary body (Haddad-Tóvolli
et al. 2015).
In the near future, it would be interesting to see progress in at least these four areas:
the role of the prechordal plate in hypothalamic specification and patterning, the
origin of hypothalamic regions, the relation between the Shh pathway and
neuroepithelial cilia and the genomic networks regulating hypothalamic neuron
specification.
Key References
Geng et al. (2008)—A network of essential regulators controls the partition of the
telencephalon into left and right hemispheres. As a consequence, in certain
genetic backgrounds, neural Shh is required for this process to be successful.
Haddad-Tóvolli et al. (2015)—Analysis of mouse mutant phenotypes shows how
different combinations of Gli transcription factors, acting downstream Shh sig-
nalling, are involved in the specification of the four classical hypothalamic
regions.
Manning L et al. (2006)—The mechanism behind the impressive change in the
pattern of Shh expression in the hypothalamus, and its role, explained.
Szabó et al. (2009)—Thirteen years after the full Shh null mutant (Chiang et al.
1996), the role of neural Shh vs prechordal Shh in hypothalamic development
were dissected by using conditional mouse mutants.
Shimogori et al. (2010)—A second conditional mutant analyzing the role of Shh
from different sources (neural or prechordal) on the developing hypothalamus.
Zhao et al. (2012)—The authors use a novel conditional mutant to show that neural
Shh is essential for the proper development of the eye and hypophysis, a role that
had escaped previous analyses.
References
Acampora D, Postiglione MP, Avantaggiato V, Di Bonito M, Vaccarino FM, Michaud J, Simeone
A (1999) Progressive impairment of developing neuroendocrine cell lineages in the hypothala-
mus of mice lacking the Orthopedia gene. Genes Dev 13:2787–2800
Altman J, Bayer SA (1986) The development of the rat hypothalamus. Adv Anat Embryol Cell Biol
100:1–178
Alvarez-Bolado G (2018) Development of neuroendocrine neurons in the mammalian hypothala-
mus. Cell Tissue Res. https://doi.org/10.1007/s00441-018-2859-1
Alvarez-Bolado G, Paul FA, Blaess S (2012) Sonic hedgehog lineage in the mouse hypothalamus:
from progenitor domains to hypothalamic regions. Neural Dev 7:4. https://doi.org/10.1186/
1749-8104-7-4
Andreu-Cervera A, Anselme I, Karam A, Laclef C, Catala M, Schneider-Maunoury S (2019) The
ciliopathy gene Ftm/Rpgrip1l controls mouse forebrain patterning via region-specific modula-
tion of Hedgehog/Gli signaling. J Neurosci 39(13):2398–2415. https://doi.org/10.1523/J.
NEUROSCI.2199-18
Aoto K, Shikata Y, Imai H, Matsumaru D, Tokunaga T, Shioda S, Yamada G, Motoyama J (2009)
Mouse Shh is required for prechordal plate maintenance during brain and craniofacial morpho-
genesis. Dev Biol 327:106–120. https://doi.org/10.1016/j.ydbio.2008.11.022
Blaess S, Szabó N, Haddad-Tovolli R, Zhou X, Alvarez-Bolado G (2014) Sonic hedgehog
signaling in the development of the mouse hypothalamus. Front Neuroanat 8:156. https://doi.
org/10.3389/fnana.2014.00156
Braun MM, Etheridge A, Bernard A, Robertson CP, Roelink H (2003) Wnt signaling is required at
distinct stages of development for the induction of the posterior forebrain. Development
130:5579–5587
Briscoe J, Thérond PP (2013) The mechanisms of Hedgehog signalling and its roles in development
and disease. Nat Rev Mol Cell Biol 14:416–429. https://doi.org/10.1038/nrm3598
Carreno G, Apps JR, Lodge EJ, Panousopoulos L, Haston S, Gonzalez-Meljem JM, Hahn H,
Andoniadou CL, Martinez-Barbera JP (2017) Hypothalamic sonic hedgehog is required for cell
2 Sonic hedgehog in Hypothalamus Development 63
Hui CC, Joyner AL (1993) A mouse model of greig cephalopolysyndactyly syndrome: the extra-
toesJ mutation contains an intragenic deletion of the Gli3 gene. Nat Genet 3:241–246
Hui CC, Slusarski D, Platt KA, Holmgren R, Joyner AL (1994) Expression of three mouse
homologs of the Drosophila segment polarity gene cubitus interruptus, Gli, Gli-2, and Gli-3,
in ectoderm- and mesoderm-derived tissues suggests multiple roles during postimplantation
development. Dev Biol 162:402–413. https://doi.org/10.1006/dbio.1994.1097
Ishibashi M, McMahon AP (2002) A sonic hedgehog-dependent signaling relay regulates growth of
diencephalic and mesencephalic primordia in the early mouse embryo. Development
129:4807–4819
Jeong Y, El-Jaick K, Roessler E, Muenke M, Epstein DJ (2006) A functional screen for sonic
hedgehog regulatory elements across a 1 Mb interval identifies long-range ventral forebrain
enhancers. Development 133:761–772. https://doi.org/10.1242/dev.02239
Jeong Y, Leskow FC, El-Jaick K, Roessler E, Muenke M, Yocum A, Dubourg C, Li X, Geng X,
Oliver G, Epstein DJ (2008) Regulation of a remote Shh forebrain enhancer by the Six3
homeoprotein. Nat Genet 40:1348–1353. https://doi.org/10.1038/ng.230
Kapsimali M, Caneparo L, Houart C, Wilson SW (2004) Inhibition of Wnt/Axin/beta-catenin
pathway activity promotes ventral CNS midline tissue to adopt hypothalamic rather than
floorplate identity. Development 131:5923–5933
Kawaguchi D, Sahara S, Zembrzycki A, O’Leary DDM (2016) Generation and analysis of an
improved Foxg1-IRES-Cre driver mouse line. Dev Biol 412:139–147. https://doi.org/10.1016/j.
ydbio.2016.02.011
Kimura S, Hara Y, Pineau T, Fernandez-Salguero P, Fox CH, Ward JM, Gonzalez FJ (1996) The
T/ebp null mouse: thyroid-specific enhancer-binding protein is essential for the organogenesis
of the thyroid, lung, ventral forebrain, and pituitary. Genes Dev 10:60–69
Kobayashi D, Kobayashi M, Matsumoto K, Ogura T, Nakafuku M, Shimamura K (2002) Early
subdivisions in the neural plate define distinct competence for inductive signals. Development
129:83–93
Krauss S, Concordet JP, Ingham PW (1993) A functionally conserved homolog of the Drosophila
segment polarity gene hh is expressed in tissues with polarizing activity in zebrafish embryos.
Cell 75:1431–1444. https://doi.org/10.1016/0092-8674(93)90628-4
Kremnyov S, Henningfeld K, Viebahn C, Tsikolia N (2018) Divergent axial morphogenesis and
early Shh expression in vertebrate prospective floor plate. EvoDevo 9:4. https://doi.org/10.1186/
s13227-017-0090-x
Lagutin OV, Zhu CC, Kobayashi D, Topczewski J, Shimamura K, Puelles L, Russell HR,
McKinnon PJ, Solnica-Krezel L, Oliver G (2003) Six3 repression of Wnt signaling in the
anterior neuroectoderm is essential for vertebrate forebrain development. Genes Dev
17:368–379
Le Gros Clark WE (1938) Morphological aspects of the hypothalamus. In: Le Gros Clark WE,
Beattie J, Riddoch G, Dott NM (eds) The hypothalamus. Morphological, functional, clinical and
surgical aspects. Oliver and Boyd, Edinburgh
Legué E, Joyner AL (2010) Genetic fate mapping using site-specific recombinases. In: Wassarman
PM, Soriano PM (eds) Guide to techniques in mouse development, Part B: Mouse molecular
genetics, 2nd edn. Elsevier, Amsterdam
Lu F, Kar D, Gruenig N, Zhang ZW, Cousins N, Rodgers HM, Swindell EC, Jamrich M,
Schuurmans C, Mathers PH, Kurrasch DM (2013) Rax is a selector gene for mediobasal
hypothalamic cell types. J Neurosci 33:259–272. https://doi.org/10.1523/JNEUROSCI.0913-
12.2013
Machold R, Hayashi S, Rutlin M, Muzumdar MD, Nery S, Corbin JG, Gritli-Linde A, Dellovade T,
Porter JA, Rubin LL, Dudek H, McMahon AP, Fishell G (2003) Sonic hedgehog is required for
progenitor cell maintenance in telencephalic stem cell niches. Neuron 39:937–950
Manning L, Ohyama K, Saeger B, Hatano O, Wilson SA, Logan M, Placzek M (2006) Regional
morphogenesis in the hypothalamus: a BMP-Tbx2 pathway coordinates fate and proliferation
through Shh downregulation. Dev Cell 11:873–885
2 Sonic hedgehog in Hypothalamus Development 65
Marin O, Anderson SA, Rubenstein JL (2000) Origin and molecular specification of striatal
interneurons. J Neurosci 20:6063–6076
Mathers PH, Grinberg A, Mahon KA, Jamrich M (1997) The Rx homeobox gene is essential for
vertebrate eye development. Nature 387:603–607. https://doi.org/10.1038/42475
Merchán P, Bribián A, Sánchez-Camacho C, Lezameta M, Bovolenta P, de Castro F (2007) Sonic
hedgehog promotes the migration and proliferation of optic nerve oligodendrocyte precursors.
Mol Cell Neurosci 36:355–368. https://doi.org/10.1016/j.mcn.2007.07.012
Miranda-Angulo AL, Byerly MS, Mesa J, Wang H, Blackshaw S (2014) Rax regulates hypotha-
lamic tanycyte differentiation and barrier function in mice. J Comp Neurol 522:876–899. https://
doi.org/10.1002/cne.23451
Mo R, Freer AM, Zinyk DL, Crackower MA, Michaud J, Heng HH, Chik KW, Shi XM, Tsui LC,
Cheng SH, Joyner AL, Hui C (1997) Specific and redundant functions of Gli2 and Gli3 zinc
finger genes in skeletal patterning and development. Development 124:113–123
Muthu V, Eachus H, Ellis P, Brown S, Placzek M (2016) Rx3 and Shh direct anisotropic growth and
specification in the zebrafish tuberal/anterior hypothalamus. Development 143:2651–2663.
https://doi.org/10.1242/dev.138305
Nüsslein-Volhard C, Wieschaus E (1980) Mutations affecting segment number and polarity in
Drosophila. Nature 287:795–801
Ono K, Yoshii K, Tominaga H, Gotoh H, Nomura T, Takebayashi H, Ikenaka K (2017) Oligoden-
drocyte precursor cells in the mouse optic nerve originate in the preoptic area. Brain Struct Funct
222:2441–2448. https://doi.org/10.1007/s00429-017-1394-2
Orquera DP, Nasif S, Low MJ, Rubinstein M, de Souza FSJ (2016) Essential function of the
transcription factor Rax in the early patterning of the mammalian hypothalamus. Dev Biol
416:212–224. https://doi.org/10.1016/j.ydbio.2016.05.021
Pak T, Yoo S, Miranda-Angulo AL, Wang H, Blackshaw S (2014) Rax-CreERT2 knock-in mice: a
tool for selective and conditional gene deletion in progenitor cells and radial glia of the retina
and hypothalamus. PLoS One 9(4):e90381. https://doi.org/10.1371/journal.pone.0090381
Patten I, Kulesa P, Shen MM, Fraser S, Placzek M (2003) Distinct modes of floor plate induction in
the chick embryo. Development 130:4809–4821. https://doi.org/10.1242/dev.00694
Pearson CA, Ohyama K, Manning L, Aghamohammadzadeh S, Sang H, Placzek M (2011)
FGF-dependent midline-derived progenitor cells in hypothalamic infundibular development.
Development 138:2613–2624. https://doi.org/10.1242/dev.062794
Placzek M, Briscoe J (2005) The floor plate: multiple cells, multiple signals. Nat Rev Neurosci
6:230–240
Puelles L, Martínez S, Martínez-de-la-Torre M, Rubenstein JLR (2004) Gene maps and related
histogenetic domains in the forebrain and midbrain. In: Paxinos G (ed) The rat nervous system,
3rd edn. Academic, San Diego, CA
Puelles L, Martinez-de-la-Torre M, Bardet S, Rubenstein JLR (2012) Hypothalamus. In: Watson C,
Paxinos G, Puelles L (eds) The mouse nervous system. Elsevier-Academic, San Diego
Rallu M, Machold R, Gaiano N, Corbin JG, McMahon AP, Fishell G (2002) Dorsoventral
patterning is established in the telencephalon of mutants lacking both Gli3 and Hedgehog
signaling. Development 129:4963–4974
Riddle RD, Johnson RL, Laufer E, Tabin C (1993) Sonic hedgehog mediates the polarizing activity
of the ZPA. Cell 75:14011416. https://doi.org/10.1016/0092-8674(93)90626-2
Roelink H, Augsburger A, Heemskerk J, Korzh V, Norlin S, Ruiz i Altaba A, Tanabe Y, Placzek M,
Edlund T, Jessell TM, Dodd J (1994) Floor plate and motor neuron induction by vhh-1, a
vertebrate homolog of hedgehog expressed by the notochord. Cell 16:761–775. https://doi.org/
10.1016/0092-8674(94)90514-2
Roessler E, Muenke M (2010) The molecular genetics of holoprosencephaly. Am J Med Genet C
Semin Med Genet 154C:52–61. https://doi.org/10.1002/ajmg.c.30236
Ruiz i Altaba A, Jessell TM (1993) Midline cells and the organization of the vertebrate neuraxis.
Curr Opin Genet Dev. 3:633–640. https://doi.org/10.1016/0959-437X(93)90100-4
66 G. Alvarez-Bolado
Abstract
Xenopus laevis (African clawed frog) is a model organism for the study of early
embryonic development of vertebrates. It is especially suited to analyze the
neuroendocrine system because it possesses hormonal systems and mechanisms
of gene regulation comparable to those in rodents and humans. The hypothalamus
is the structure of the forebrain that plays a fundamental role in the neuroendo-
crine and homeostatic regulation of the body, representing a central hub for the
neural networks involved in these functions. As such the hypothalamus is the
nexus between the autonomic and limbic systems, which condition the behavioral
and regulatory responses of the animal. During the development of Xenopus, the
hypothalamus is located rostral to the diencephalon, possesses derivatives of the
alar and basal plates, and includes molecular compartments, comparable to
amniotes (reptiles, birds, and mammals), expressing transcription factors such
as Nkx2.1, Islet 1, and Otp in unique subpopulations.
Keywords
Prosencephalon · Forebrain patterning · Expression patterns · Evolution
Abbreviations
BH Basal hypothalamus
CT Caudal tuberal region
Dien Diencephalon
hp1 Hypothalamic prosomeric domain 1
hp2 Hypothalamic prosomeric domain 2
Hyp Hypophysis
M Mammillary region
Ma Mammillary area proper
oc Optic chiasm
P1 Prosomere 1
P2 Prosomere 2
P3 Prosomere 3
P3b Basal plate of P3
Pa Paraventricular nucleus
PA Pallium
Pac Caudal paraventricular nucleus
Par Rostral paraventricular nucleus
PO Preoptic region
PT Pretectum
PTh Prethalamus
RM Retromammillary region
RT Rostral tuberal region
Sag Sagittal
SC Suprachiasmatic region
SCc Caudal suprachiasmatic region
SCr Rostral suprachiasmatic region
SPa Subparaventricular area
SPA Subpallium
Th Thalamus
Tub Tuberal region
From the late 1940s to the 1970s, when modern pregnancy tests were not available,
Xenopus laevis frogs were injected with the urine of possibly pregnant women and a
positive result was indicated when the presence of human chorionic gonadotropin
(hCG) induced the frogs to lay a large number of eggs (Lee-Liu et al. 2017). At the
same time, in many laboratories, this ease with which female frogs could be induced
to lay eggs, which could subsequently be fertilized in vitro, along with the relatively
large size of the eggs and embryos, and their development ex utero, made Xenopus
laevis an excellent model organism to study the early embryonic development of
vertebrates. In addition, due to its metamorphosis phase, this model has provided
insight into the developing neuroendocrine systems, as well as a model to study
endocrine disrupters. In fact, in recent years evidence has accumulated that persistent
organic pollutants, heavy metals, pharmaceutical and personal care products, agri-
cultural chemicals, and aerospace products are acting as endocrine disrupters that
may alter the activity of hypothalamus–pituitary–thyroid axis and subsequently
3 Development of the Hypothalamus in Xenopus laevis 69
interfere with the endocrine systems of both wildlife and humans by affecting
reproductive biology and the thyroid system. In this context, it was also
demonstrated that hormonal control of development during the human perinatal
period is critically important and complex, with multiple hormones that regulate
fetal growth, brain development, and organ maturation in preparation for birth.
Notably, many significant parallels exist between metamorphosis of frogs and the
birth of mammals, and underlying these analogous developmental changes are
conserved neuroendocrine components (Buchholz 2017). Frogs and mammals
have identical hormones that reach their peak at metamorphosis and birth, have
conserved hormone receptors and mechanisms of gene regulation, and share com-
parable roles for hormones in many target organs. Indeed, in both frogs and
mammals, genetic and environmental perturbations of hormonal control can cause
irreversible morphological and physiological changes and can predispose
individuals to diseases of adulthood.
The hypothalamic development in Xenopus has been the focus of thorough
studies in recent years (Domínguez et al. 2013, 2014, 2015). In these studies,
Xenopus was selected mainly due to the convenience of “hormone induced” breed-
ing that makes embryos available for experimentation at all times of the year and that
they are easy to maintain under laboratory conditions. In addition, an accurate
timetable of development exists (Nieuwkoop and Faber 1967) that allows compara-
tive interpretation of developmental processes. Notably, analysis of the amphibian
hypothalamus serves as a model for understanding perinatal events in mammals due
to the analogies in development of these structures.
In all vertebrates, the hypothalamus plays a key role in survival of the animal, acting
as the bridge between the external (visual, olfactory, and gustatory) and internal
stimuli and endocrine, autonomic and limbic responses. Thus, the hypothalamus is
considered the forebrain territory par excellence for controlling homeostatic pro-
cesses. The hypothalamus has been studied to a great extent, however the interpre-
tation of its regionalization during development has been debated throughout the last
decades and additional studies may shed further light on this fundamental,
anatomical question.
From an anatomical point of view, the strong curvature of the neural tube flexure
at the junction of the midbrain and forebrain forms a sharp cephalic flexure involving
the hypothalamus (see Fig. 3.1). This developmental event has made it hard to
identify the basic units or subdivisions in the mature hypothalamus and understand
the topological relationships between them. Thus, following mere topographical
positions and interpreting “classical” transverse sections, the hypothalamus was
considered a diencephalic region beneath the thalamus and different nomenclatures
have been used for nuclei and hypothalamic regions in different vertebrates. The
70 N. Moreno and A. González
Fig. 3.1 Schematic comparison of the hypothalamic development between mammals and
non-mammals. At the neural plate stage (a), the brain undergoes neurulation forming the neural
tube (b). The hypothalamic anatomy is influenced by the flexure of the neural tube. In mammals the
longitudinal axis bends almost 90 forming a sharp flexure and the rostral tube is moved to a
“ventral” position, whereas in non-mammals this angle is less pronounced. Additionally, the
telencephalic evagination in the case of mammals is huge and proportional over hp1 and hp2,
flattening the hypothalamic territory, whereas in non-mammals, this event pushes hp2, turning the
hypothalamus more “ventral.” In contrast to non-mammals, the hypothalamic evagination in
mammals contributes to the elongation of the hypothalamic territory
that are also defined histogenetic domains based on regulatory gene expression
patterns. According to the original prosomeric model and its subsequent revisions,
the hypothalamus is excluded from the diencephalon, which is composed of three
neuromeres (prosomeres P1–P3, Fig. 3.1). The rostral-most forebrain is designated
the secondary prosencephalon that contains the hypothalamus (rostral to
diencephalic P3), the telencephalon impar, and the telencephalic hemispheres
(Puelles and Rubenstein 2003). The interpretation of the parts of the secondary
prosencephalon is fraught with difficulties, mainly derived from the early optic and
hemispheric evaginations and the different degree of development shown across
vertebrates that disturb the primary pattern of this region. However, morphological,
molecular, and hodological (connectivity) results have continued to highlight the
organization of the main parts of the secondary prosencephalon and its subdivisions,
particularly in mice, although comparable results have been obtained in other
amniotes, amphibians, and fishes (reviewed in Domínguez et al. 2015; Ferran
et al. 2015).
Based on gene expression patterns and other molecular markers, interpreted
under the current prosomeric model (Fig. 3.1), such as FoxG1, the preoptic region
(PO, see Fig. 3.2) is part of the subpallium of the telencephalon (SPA, Fig. 3.2),
instead of the hypothalamus, where it was traditionally included (Puelles et al. 2012).
It forms the dorsal boundary of the hypothalamic territory. In the proper hypothala-
mus, a transverse and complete limit, called intrahypothalamic boundary, has been
described marking anteroposterior patterning and defining subdivisions. Thus, ter-
minal (rostral, hp2: hypothalamic prosomeric domain 2, Fig. 3.1 dark blue area) and
peduncular (caudal, hp1: hypothalamic prosomeric domain 1, Fig. 3.1 lighter blue
area) regions were distinct on the basis of the differential expression of transcription
factors in the early neural tube. In addition, a band of Nkx2.2 expressing cells
becomes bent at the cephalic flexure and highlights the continuity of the dorsoventral
patterning across the hypothalamus, revealing the existence of the longitudinal
organization in the forebrain, and indicating alar versus basal territories. As such,
four main domains can be considered in the hypothalamus, because both the hp1 and
hp2 segments contain alar and basal parts (see Fig. 3.1). Furthermore, two longitu-
dinal subdivisions are also formed in the alar and basal hypothalamic regions. The
paraventricular area (Pa), topologically the dorsalmost part of the hypothalamus, is
located adjacent to the preoptic region and mainly expresses the transcription factors
Tbr1, Sim1, and Otp, and produces the magnocellular and parvocellular neurosecre-
tory cell populations of the paraventricular, supraoptic, and periventricular nuclei.
The second domain is the subparaventricular area (SPa), which forms the ventral part
of the alar hypothalamus and is characterized by the expression of Dlx genes and
gives rise to the suprachiasmatic, the anterior hypothalamic nucleus, and the
subparaventricular zone. Sonic hedgehog and the transcription factor Nkx2.1
specifically characterize the basal hypothalamus formed by two longitudinal
subdivisions: tuberal (Tub) and mammillary (M) regions. The tuberal region
produces the anterobasal and posterobasal areas, and the ventromedial, dorsomedial,
and arcuate nuclei. Rostrally to all those areas, in the terminal region a distinct
acroterminal territory has been defined (see Fig. 3.1, yellow dots, Ferran et al. 2015).
72 N. Moreno and A. González
Fig. 3.2 (1) Schematic representation of the timing of the main developmental events during
Xenopus laevis development (based on Nieuwkoop and Faber 1967) pf: post-fertilization. (2) Brain
region changes from embryonic to larval. Photomicrographs of Nissl-stained sagittal (b, h) and
transverse (c–f; i–l) sections through the hypothalamus of an embryonic (b–f) and a larval (h–l)
Xenopus; from anterior to posterior, at the approximate levels indicated on the schemes (a, g). These
photomicrographs illustrate the segregated regions recognized in the hypothalamus using
techniques of the time (Nieuwkoop and Faber 1967). The scale bar in c is valid for b–
f ¼ 100 μm. The scale bar in i is valid for h–l ¼ 200μm
It corresponds to the rostral-most portion of the neural tube and is responsible for the
development of structures such as the lamina terminalis and the optic chiasm in the
alar hypothalamus, whereas in the basal hypothalamus it gives rise to the median
eminence and the associated arcuate nucleus, the infundibulum, and the neurohy-
pophysis (Morales-Delgado et al. 2014; Puelles et al. 2012).
development of the Xenopus brain, including the hypothalamus, has been the basis
of numerous studies using a variety of experimental techniques and protocols. As
with any laboratory model organism, Xenopus possesses standard protocols for
laboratory breeding and husbandry, techniques to modify gene expression (e.g.,
antisense morpholino oligonucleotides), the generation of transgenic animals, and
genome editing (e.g., CRISPR/Cas9 system; Bhattacharya et al. 2015). In addition,
there are currently genetic and genomic data available from Xenopus to analyze
genes, gene families, and gene networks (see Tandon et al. 2017).
Throughout development, brain changes can easily be correlated with the striking
body transformations that take place through embryonic and larval development.
These include changes in the hypothalamus that progressively lead to specific
developmental events related to hormonal control, sexual maturity and behavioral
events which functionally are related to maturation of the hypothalamus.
74 N. Moreno and A. González
For Xenopus laevis, Nieuwkoop and Faber (1967) described and illustrated in
precise tables, the different stages of embryonic and larval development mainly
based on external anatomy. This staging is still used today (Fig. 3.2, 1). Xenopus
laevis development encompasses embryos (stages 1–45) and larvae (stages 46–66),
at stage 66 metamorphosis is completed and the tadpole becomes a froglet.
According to these tables, from stage 24 the animal begins to show initial motor
reactions to external stimulation, and by stage 26 such movements are spontaneous.
But it is only from stage 28 that changes in brain morphology related to the
hypothalamus are detected, with the hypophyseal primordium (future pituitary)
penetrating beyond the optic stalk, and the infundibulum rapidly developing by
progressive thinning of the ventral wall. From stage 29/30, the chiasmatic ridge is
rostrally separated from the lamina terminalis, and it is just later, at stage 31, that the
hypophyseal cell plate reaches the caudal border of the chiasmatic ridge, but it will
be connected to the ectoderm until stage 32. All of these events occur at early
embryonic developmental stages. Thus, during late embryonic stages, the hypothal-
amus is already present as an anatomical entity, but its maturity is likely linked to
subsequent changes in the animal as a whole. The next main event that takes place
with respect to behavioral responses is the beginning of independent feeding at stage
45, when larval stages begin. At stage 47, the pars intermedia of the hypophysis
begins to differentiate, and at the same stage, the genital ridges begin to protrude into
the coelomic cavity. At stage 49 the infundibular extension of the neurohypophysis
begins to develop and the first anlage of tuberal part of the adenohypophysis can be
observed. From stage 50, the gonadal rudiments form and the thyroid follicles can be
observed, but it is not until stage 52 that sexual differentiation of the gonads begins,
and at stage 55 the first meiotic divisions are observed.
Thus, in Xenopus, the development of the hypothalamus actually takes place over
a relatively long period, throughout embryonic and larval development (Fig. 3.2, 2).
This makes this system a good model to dissect apart the signals that lead to a
mature, functional hypothalamus. Of note, it is known that from the cell blastomere
of stage 16, hypothalamic dopaminergic cells can already be detected (Huang and
Moody 1992). The development of the anterior hypothalamus occurs mainly around
embryonic stages 20–25, while the posterior part develops later, through the late
embryonic and early larval stages 40–52. In contrast, the preoptic area which is
mainly related to reproductive and parental behaviors, and now defined as part of the
telencephalon, develops almost at the end of development, from stage 58 (Tay and
Straznicky 1982).
3 Development of the Hypothalamus in Xenopus laevis 75
3.2.2.2 Neuropeptides
Neurochemically different cell populations in the Pa of the hypothalamus in Xenopus
include neurons containing several neuropeptides such as somatostatin (SOM;
Fig. 3.3k), mesotocin (MST; Fig. 3.3j; an oxytocin-like peptide), arginine-vasotocin
(AVT), corticotropin-releasing hormone (CRH), thyrotropin-releasing hormone
(TRH), or orexins (González et al. 1995; Domínguez et al. 2015; López et al.
2017). These cell populations project to the pars nervosa of the pituitary gland
(Tuinhof et al. 1994), and a profuse connection exists between this region and the
tuberal hypothalamus (Tub, Figs. 3.1 and 3.4; Domínguez et al. 2013) that may be
involved in regulating ingestion of food. Numerous opsin-positive neurons have
been found in the previously named “hypothalamic magnocellular preoptic
nucleus,” and most of them co-express MST. Axonal processes from these cells
reach the neural lobe of the pituitary gland, suggesting that in Xenopus,
mesotocinergic hypothalamic neurons, which are directly involved in photorecep-
tion and could be regulating seasonal adaptation changes as well. (Álvarez-Viejo
et al. 2003). Independent of function, the early emergence of cells expressing TRH
or MST and SOM, and the presence of Otp has been correlated (Fig. 3.3j, k;
Domínguez et al. 2013, 2015), suggesting that the transcription factor Otp may
participate in the acquisition of their phenotypes. Thus, in both Xenopus and mouse,
transcription factors are involved in the specification of hypothalamic postmitotic
populations and in the differentiation of independent nuclei within this territory.
76 N. Moreno and A. González
Fig. 3.3 Alar hypothalamic regions identified by molecular markers. Upper panel:
Photomicrographs of sagittal (c, e–h, l, m) and transverse sections (a–d0 , i–k, n–g) through the
developing Xenopus alar hypothalamic region showing the single expressions of Otp (a), xShh
(e, n), and Nkx2.1 (f, o), and the combined expressions of Isl1 and Nkx2.1 (g), Isl1/Otp (b, c, h),
Nkx2.2/Otp (d, i), Nkx2.2/MST (j), Nkx2.2/SOM (k), xDll4/Otp (l), Nkx2.1/Nkx2.2 (m), xShh/TH
(p), and Nkx2.1/TH (q). The arrowheads in d0 and i indicate double labeled cells, and in (a) indicate
Otp ventricular expressing cells. The yellow line in c indicates the level of the section shown in c0 .
Developmental stages are indicated in each photograph. Lower panel: Schematic representation (r)
summarizes the combinatorial code of markers present in the alar hypothalamus and its adjacent
domains in lateral and transverse views (r). MST, mesotocin; TH, tyrosine hydroxylase. Scale bars
in a, b, e, h, j–m ¼ 50 μm; in c, d, g, i, n–q ¼ 100 μm. All photomicrographs are modified from
Domínguez et al. (2013)
3 Development of the Hypothalamus in Xenopus laevis 77
Fig. 3.4 Basal hypothalamic regions identified by molecular markers. Upper panel:
Photomicrographs: transverse (c, e, g, h, i, k–p) and sagittal (a, b, d, f, j, q) sections through the
developing Xenopus basal hypothalamic region showing the single expressions of xShh (a) and the
combined expressions of Isl1 and Nkx2.1 (b, c), Isl1 and Otp (d, e), Nkx2.2 and Otp (f, g), Pax7 and
Nkx2.1 (h–k), Pax7 and Otp (l), TH and Nkx2.1 (m), Shh (n), Isl1 (o), Nkx2.2 (p, q). The
arrowheads in h, k, and m indicate double labeled cells. The red line in q indicates the alar/basal
boundary. Developmental stages (st) are indicated in each photograph. Lower panel: schematic
representation summarizes the combinatorial code of markers present in the basal hypothalamus
and its adjacent domains in lateral and transverse views. TH: tyrosine hydroxylase. Scale bars in a–
i; k–q ¼ 100 μm; in j ¼ 50 μm. All the photomicrographs are modified from Domínguez et al.
(2014)
78 N. Moreno and A. González
likely originate from the adjacent Pax7 expressing population of the P3 basal plate
(Fig. 3.4h–l), which indicates that part of the mammillary region arises in the
prethalamic tegmentum. Fate map analysis is needed to confirm this. In Xenopus,
the Lhx7 expression in the alar hypothalamus continues ventrally reaching the
mammillary territory (Domínguez et al. 2013), suggesting the existence of a
tuberomammillary terminal zone in anurans as in mammals (Domínguez et al.
2014). The presence of a catecholaminergic cell population (TH positive, González
et al. 1994; Domínguez et al. 2014) that coexpresses Nkx2.1 and Otp (Fig. 3.4n–q)
also characterizes the mammillary region. As for the specification of this cell
population, there is currently controversy about its origin, but several data support
a contribution of diencephalic areas to the mammillary territory since Pax7
expressing cells likely originated in P3 and subsequently colonized the mammillary
region (Domínguez et al. 2014).
In multiple species analyzed in recent years, it has been described that the hypothal-
amus includes comparable molecular compartments, and each compartment shows a
tendency to a common organization with respect to the profile of molecular expres-
sion, but at the same time, the hypothalamic organization is subject to evolutive/
adaptive changes/modifications during the anamnio-amniotic transition. In this
context, the evolutionary leap of Xenopus, an amphibian (amphibia are the only
group of anamniote tetrapods), to the rest of the tetrapods, implies relevant adaptive
changes for the conquest of the water-independence of the new environment, and
therefore it has some obvious consequences in the organization of the brain. The
main common features of the hypothalamus are its position and regionalization,
since in all models it is topographically located rostral to the diencephalon and it is
divided into alar and basal regions. In addition, each of them is rostro-caudally
subdivided into two different domains according to molecular criteria (reviewed
in Domínguez et al. 2015). Moreover, the chemoarchitectonic profiles of the
territories functionally define them by the presence of multiple postmitotic cell
populations occupying neuroendocrine territories. In contrast, in amniotes a
preoptohypothalamic boundary is evident, which is likely related to the higher
complexity of the paraventricular area that starts to express transcription factors
such as Pax6. In addition, in this alar territory, the restricted expression of Nkx2.1 in
the subparaventricular region observed in Xenopus is gradually more restricted
reaching a total lack in mammals, which could be related to the gradual pallial and
thalamic expansion that take place in amniotes.
3 Development of the Hypothalamus in Xenopus laevis 81
Key References
References
Álvarez-Viejo M, Cernuda-Cernuda R, DeGrip WJ, Álvarez-Lopez C, García-Fernández JM (2003)
Co-localization of mesotocin and opsin immunoreactivity in the hypothalamic preoptic nucleus
of Xenopus laevis. Brain Res 969(1–2):36–43. https://doi.org/10.1016/S0006-8993(03)02273-X
Bhattacharya D, Marfo CA, Li D, Lane M, Khokha MK (2015) CRISPR/Cas9: an inexpensive,
efficient loss of function tool to screen human disease genes in Xenopus. Dev Biol 408
(2):196–204. https://doi.org/10.1016/j.ydbio.2015.11.003
Buchholz DR (2017) Xenopus metamorphosis as a model to study thyroid hormone receptor
function during vertebrate developmental transitions. Mol Cell Endocrinol 459:64–70. https://
doi.org/10.1016/j.mce.2017.03.020
Davis AM, Seney ML, Walker HJ, Tobet SA (2004) Differential colocalization of Islet-1 and
estrogen receptor alpha in the murine preoptic area and hypothalamus during development.
Endocrinology 145(1):360–366. https://doi.org/10.1210/en.2003-0996
Domínguez L, Morona R, González A, Moreno N (2013) Characterization of the hypothalamus of
Xenopus laevis during development. I. The alar regions. J Comp Neurol 521(4):725–759.
https://doi.org/10.1002/cne.23222
Domínguez L, González A, Moreno N (2014) Characterization of the hypothalamus of Xenopus
laevis during development. II. The basal regions. J Comp Neurol 522(5):1102–1131. https://doi.
org/10.1002/cne.23471
Domínguez L, González A, Moreno N (2015) Patterns of hypothalamic regionalization in
amphibians and reptiles: common traits revealed by a genoarchitectonic approach. Front
Neuroanat 9:3. https://doi.org/10.3389/fnana.2015.00003
Ferran JL, Puelles L, Rubenstein JL (2015) Molecular codes defining rostrocaudal domains in the
embryonic mouse hypothalamus. Front Neuroanat 9:46. https://doi.org/10.3389/fnana.2015.
00046
González A, Marin O, Tuinhof R, Smeets WJ (1994) Ontogeny of catecholamine systems in the
central nervous system of anuran amphibians: an immunohistochemical study with antibodies
82 N. Moreno and A. González
Abstract
Neurons in the hypothalamic arcuate nucleus (ARC) relay and translate important
cues from the periphery into the central nervous system to maintain critical
homeostatic processes such as energy balance, growth, and reproductive
behaviors. Despite extensive studies dedicated to unraveling the physiological
functions of various neuronal types in the ARC, the gene regulatory networks that
drive their development are still poorly understood. This chapter provides a
concise overview of the transcription factors that control the development of
ARC neurons, followed by discussions of several outstanding challenges in the
field and future directions to address those issues using the recent technological
advances in genome-wide studies.
Keywords
Arcuate nucleus · Gene regulation · Transcription factor · POMC · AgRP · NPY ·
GHRH
J. W. Lee (*)
Department of Biologial Sciences, University at Buffalo, Buffalo, NY, USA
e-mail: [email protected]
C. Huisman
Department of Pediatrics, Oregon Health & Science University, Portland, OR, USA
S. Lee
College of Pharmacy, Research Institute of Pharmaceutical Sciences, Seoul National University,
Seoul, Korea
4.1 Introduction
Fig. 4.1 Different neurons populate the arcuate nucleus (ARC). The ARC contains as many as
24 neuronal cell types according to the recent scRNA-seq results (Campbell et al. 2017). Only a few
types of neurons have been extensively characterized, including tuberoinfundibular dopaminergic
(TIDA)-, agouti-related peptide (AgRP)-, pro-opiomelanocortin (POMC)-, growth hormone
relasing hormone (GHRH)- and Kiss1-neurons. GH growth hormone, GnRH gonadotropin releas-
ing hormone
4 Gene Regulatory Programs in the Development of Hypothalamic Arcuate. . . 85
neurons (Campbell et al. 2017). Specific roles for each of these ARC neuronal
populations are just beginning to be addressed. For instance, in female mice,
AgRP-neurons that co-express somatostatin (Sst) were shown to specifically mediate
corticotropin-releasing factor (CRF)-driven thermogenesis to cope with environ-
mental challenges (Kuperman et al. 2016). Unfortunately, the gene regulatory
networks that direct the development of arcuate neurons, which are the subject of
this chapter, are poorly understood. Clearly, this information is needed to enable our
understanding of the developmental trajectories through which common progenitors
differentiate to distinct types of arcuate neurons (Biebermann et al. 2012; Bluet-Pajot
et al. 2001; Hill et al. 2008; Hrabovszky 2014) and provide crucial insight into how
and when the seemingly homogeneous populations of neurons segregate to multiple,
functional subpopulations (Campbell et al. 2017).
Development of the hypothalamus involves multiple organizing signals and
transcription factors. In mice, prior to embryonic day 10 (E10), the diencephalic
region is separated from surrounding regions by the influence of organizing signals
such as Wnts (for wingless/integrins), Shh (for sonic hedgehog), Bmps (for bone
morphogenetic proteins, and Fgfs (for fibroblast growth factors). These organizing
signals eventually induce the expression of the homeodomain (HD) transcription
factor (TF) Nkx2.1, which is a key marker of hypothalamic tissue (Martinez-Ferre
and Martinez 2012). During the subsequent neurogenic period (E10–16), neural
stem cells in the ventricular zone give rise to neural precursors (Ifft 1972; Ishii and
Bouret 2012). Because neurons, astrocytes, and oligodendrocytes all derive from the
same progenitor pool, the proneural TFs that characterize the neurogenic period
simultaneously drive neurogenesis and suppress gliogenesis (Rowitch and Kriegstein
2010). Notably, neurogenesis in the hypothalamus proceeds in an outside-in fashion
(Altman and Bayer 1978). Neurons in the most lateral nuclei, such as the lateral
hypothalamus area (LHA), are generated first. This neurogenic period is followed by
birth of cells that will reside in more medially located nuclei such as the dorsomedial
nucleus (DMH) and the ventromedial nucleus (VMH), and finally by neurons in
nuclei along the midline such as the ARC (Byerly and Blackshaw 2009).
In this chapter, we give an overview of the findings related to the gene regulatory
programs in the development of various arcuate neuronal types, which have been
elucidated over the past few decades. We then highlight several outstanding
challenges in the field, together with future studies to address those issues using
the recent technological breakthroughs in genome-wide studies.
Fig. 4.2 Expression of Otp, NPY, Dlx1, GHRH in the developing arcuate nucleus (ARC). The
expression in transverse brain sections of (left) the transcription factors (TFs) Otp (red) and Dlx1
(green) detected with immuno-fluorescent histochemistry (IHC); and (right) mRNAs for
neuropeptides neuropeptide Y (NPY (a marker of agouti-related peptide (AgRP)-neurons, which
derive from Otp+ cells) and growth hormone releasing hormone (GHRH, a marker of GHRH-
neurons, which derive from Dlx1+ cells) using in situ hybridization (ISH). The ARC region is
denoted, left, as a yellow (dashed) triangle, and right, by autoradiographic sliver grains (black).
E15.5, 17.5: embryonic days of development
4 Gene Regulatory Programs in the Development of Hypothalamic Arcuate. . . 87
4.2.2 Retina and Anterior Neural Fold Homeobox (Rax, aka Rx)
The HD TF Rax, a key regulator of early retinal development (Zhang et al. 2000),
also plays crucial roles in the formation of the ventral neuroendocrine hypothalamus
(Muranishi et al. 2012). Nkx2.1-expressing neurons, which include many develop-
ing neurons in the ARC, appear to be derived from Rax-expressing cells. Elimination
of Rax throughout ARC progenitors in mice results in loss of ventral hypothalamic
expression of Nkx2.1 and POMC (Lu et al. 2013). These results suggest that Rax
controls the production of POMC and possibly other ARC neuronal cell types, given
the finding that POMC-derived progenitor cells differentiate into multiple ARC
neuronal types (Padilla et al. 2010; Sanz et al. 2015).
4.2.4 Isl1
4.2.5 Bsx
The HD TF Bsx is expressed in the ARC as early as E15.5 (Cremona et al. 2004). In
the adult mouse brain, Bsx expression is restricted to the hypothalamus with
prominent expression in the ARC and the DMH as well as some cells in the LHA
(Sakkou et al. 2007). In Bsx mutant brains, the expression of AgRP and NPY is
downregulated but not abolished (Sakkou et al. 2007). Although it is unclear if Bsx
is required for the early fate specification of AgRP-neurons, it is possible that Bsx
and Isl1 have essential but redundant roles in the fate specification of AgRP-neurons
as well as their upregulation of AgRP (Lee et al. 2016).
In the mouse (from E12.5), the HD TF Otp is expressed in neurons that give rise to
the PVN, supraoptic (SON), anterior periventricular (aPV), and ARC nuclei
(Simeone et al. 1994). Otp-null mice die soon after birth and display progressive
impairment of crucial developmental events, including reduced cell proliferation,
4 Gene Regulatory Programs in the Development of Hypothalamic Arcuate. . . 89
The HD TFs Dlx1 and Dlx2 are expressed in the region of the developing and adult
ARC (Eisenstat et al. 1999; Saino-Saito et al. 2003). Dlx1/2 plays important and
redundant roles in the fate specification of GABA+ and TH+ neurons in the forebrain
(Andrews et al. 2003; Cobos et al. 2007; Long et al. 2007; Petryniak et al. 2007; Qiu
et al. 1995). The immunochemical analyses with a pan-Dlx antibody detecting
multiple Dlx factors (including Dlx1/2) reveal that Dlx factors are also expressed
in TH+ neurons and non-AgRP-neuronal type GABA+ neurons in the ARC (Yee
et al. 2009). Adult Dlx1-null mice show a two-fold reduction in the number of TH+
neurons in the ARC (Yee et al. 2009), whereas the role of Dlx2 in the hypothalamus
remains unexplored. We recently reported that Dlx1 and Dlx2 are highly expressed
in TH+ neurons in the ARC, including GHRH-neurons and play critical roles in the
production of GHRH-neurons (Lee et al. 2018).
The bHLH TF Nhlh2 is expressed in the developing nervous system as early as E9.5
(Begley et al. 1992; Gobel et al. 1992). Nhlh2-null mice show hyperphagia and
obesity as well as infertility (Good et al. 1997). These mice show no alteration in
expression of POMC mRNA, although their αMSH levels are reduced, which is due
to a reduction in the expression and activity of prohormone convertases (Pc1/3) that
cleave POMC to adrenocorticotropic hormone (ACTH) and αMSH (Jing et al.
2004). The infertility of Nhlh2-null mice is explained by the finding that migration
of GnRH-neurons to their final hypothalamic sites is significantly reduced in Nhlh1/
Nhlh2-double KO mice (Kruger et al. 2004). Interestingly, POMC-neurons are
reduced in mice when Nhlh2 is deleted specifically in GnRH-expressing neurons
(Schmid et al. 2013), suggesting a critical role for GnRH cells in the development of
POMC-neurons.
4.2.12 Rbpjk
4.2.13 Ikaros
From the public databases such as GenePaint (Visel et al. 2004) and Allen Brain
Atlas (Jones et al. 2009; Lein et al. 2007) as well as the recent expression studies of
Shimogori et al. (2010) (Box 4.2), it is clear that more TFs are expressed in the
developing ARC, which include the TFs Pou2f2, Sox14, Tbx2, Tbx19, Arx, Pbx1,
Insm1, and Lhx5. Future investigation of the role of these TFs in arcuate neuronal
development will allow us to establish which TFs, alone or together, play crucial
roles in regulating the production of the many arcuate neuronal types during
embryogenesis.
It is important to note that what appear to be distinct classes of arcuate neurons are
indeed, functionally interconnected among themselves. For instance, reproduction, a
highly energy-demanding process, is coupled to energy balance, and leptin serves as
a key mediator of this coupling by controlling both food intake via AgRP-/POMC-
neurons and reproduction via Kiss1-neurons (Chehab 2014). In addition, AgRP-
neurons project onto GnRH and Kiss1-neurons while POMC-neurons project onto
Kiss1-neurons, forming a circuit that allows for cross talk between food intake and
reproduction (Chehab 2014). Another example is the well-known correlation
between nutritional status and linear growth. As such, GH/GHRH levels are
inversely correlated with circulating levels of the anorexic hormone leptin, which
in turn targets arcuate neurons that control energy balance (Considine et al. 1996;
Dieguez and Casanueva 1995; Grinspoon et al. 1996). Clearly, the generation of all
of the different ARC neuronal subtypes is critical for homeostasis. Currently,
evidence suggests that production of multiple arcuate neurons during embryogenesis
involves common progenitors. These progenitors then differentiate into multiple
types of neurons via gene regulatory programs that involve distinct fate-specifying
TFs. How these programs are initiated/specified is unclear, but these programs allow
concise regulation of each developmental lineage. In this context, it is interesting to
note that POMC+ precursors generate several arcuate neuronal types, including
POMC-, AgRP-, and Kiss1-neurons (Padilla et al. 2010; Sanz et al. 2015). It is
92 J. W. Lee et al.
also possible that the production of GHRH-neurons is coordinated with the genera-
tion of AgRP- and/or POMC-neurons via common progenitors.
In addition to identifying and characterizing the TFs that drive the development of
arcuate neuronal subtypes, several challenges remain to be addressed to fully
understand the gene regulatory programs that direct this developmental process. In
particular, three issues are noteworthy.
First, for most TFs that are known to control the development of arcuate neurons
(Table 4.1), information on their critical downstream target genes is largely missing.
Determining the binding sites of these TFs in the different neuronal cell types of the
ARC will help unravel the detailed molecular basis of their action in fate specifica-
tion and development of arcuate neurons. This information should also facilitate our
understanding of how different arcuate neurons coordinately develop during
embryogenesis.
Second, precise assessment of the whole spectrum of actions that each TF plays in
fate specification and development of the different arcuate neuronal types is seri-
ously hampered by the fact that we currently have only a limited number of early fate
markers for each arcuate neuronal population. Thus, we are lacking tools to identify
lineage specification pathways.
Third, directly following is that Cre lines that can delete genes of interest in a
specific population of arcuate neurons during earlier fate specification stages of
development are missing. Such Cre lines will allow us to delineate the developmen-
tal action of each TF in specific populations of arcuate neurons independent of its
action in other cell types. Potential resolutions to these issues, using recent advances
in genome-wide studies, are discussed below.
The first step to reconstruct the gene regulatory programs for arcuate neuronal
development is to identify TFs that play critical roles in arcuate neuronal develop-
ment (Table 4.1). The next step is to identify the genome-wide target genes of each
of these TFs in the developing ARC. This will also facilitate our effort to decipher
how these TFs function, alone and/or in combination, to coordinate the production of
different populations of arcuate neurons. A strategy used in our recent study provides
a template for such work. We found that the HD TFs Otp and Dlx1 (and its homolog
Dlx2) play essential roles in the development of AgRP- and GHRH-neurons,
respectively (Lee et al. 2018). In addition to a loss of GHRH-neurons, we found
that mice with Nkx2.1-Cre-directed cKO for the adjoining Dlx1 and Dlx2 genes
4 Gene Regulatory Programs in the Development of Hypothalamic Arcuate. . . 93
Fig. 4.3 The Dl1/2-Otp regulatory axis in the development of GHRH- and AgRP-neurons. (a)
Summary of deficits in the development of growth hormone releasing hormone (GHRH)- and
agouti-related peptide (AgRP)-neurons in Otp/ and Dlx1/2-conditional knockout (cKO) mice.
While Otp/ mice lose AgRP-neurons, Dlx1/2-cKO mice lose GHRH-neurons but gain AgRP-
neurons. (b) CHIP-seq schematic, including cross-linking of tissues, preparation of lysed chromatin
and sonification, immunoprecipitation of chromatin fragments bound by transcription factor
(TF) using antibody against the TF, and reverse cross-linking and isolation of DNA, followed by
library constructiuon, sequencing and analyses. (c) Our CHIP-seq experiments using anti-Dlx1
antibody as well as the ensuing functional assays revealed that Dlx1 suppresses the expression of
Otp by directly binding the silencer region located downstream of the Otp gene in developing
GHRH-neurons (Lee et al. 2018). Direct target genes of Otp in developing AgRP-neurons and of
Dlx1/2 in developing GHRH-neurons remain to be investigated.
Fig. 4.4 De novo motif analysis and partner transcription factors (TFs). (a) Unbiased search of
CHIP-seq peaks for motifs that recruit the target TF reveals multiple modes of DNA binding by the
TF. TFs often bind their target sites indirectly through other TFs (e.g., X in this diagram) or in
synergy with another TF (e.g., Y in this diagram). (b) The homeodomain (HD) TF Isl1 is known to
pair with other TFs in a cell type-specific manner during development, and this combinatorial mode
of actions allows one TF to control multiple different fates
common progenitors and that developing GHRH-neurons are still plastic with the
potential to differentiate to AgRP-neurons. Thus, suppression of Otp expression by
Dlx1/2 forces the progenitors to take the lineage to GHRH-neurons rather than the
erroneous pathway to AgRP-neurons. These results demonstrate that the Dlx1/2-Otp
gene regulatory axis directs a balanced production of GHRH- and AgRP-neurons
during embryogenesis, contributing to the postnatal homeostatic functions of hypo-
thalamic arcuate neurons that control energy balance and growth.
In addition to directly binding their cognate binding motifs, TFs often bind DNA
either indirectly through protein–protein interactions with other DNA-bound TFs
(e.g., X in Fig. 4.3a) or synergistically through adjacent motifs for other TFs (e.g., Y
in Fig. 4.4a). Once the genome-wide occupancy of a TF is determined using the
CHIP-seq technology, de novo motif analysis of CHIP-seq peaks (Bailey et al. 2009;
Machanick and Bailey 2011) will allow us to predict its different modes of DNA
binding as well as partner TFs. Hence, de novo motif analyses, coupled with the
ensuing validation experiments, also provide crucial insights into the identity of
partner TFs that function together with the original TF (e.g., X and Y in Fig. 4.4a),
contributing to our understanding of how distinct fate-determining TFs are
integrated into the gene regulatory programs for arcuate neuronal development. It
is possible that some of the TF in Table 4.1 may be engaged in such protein–protein
interactions, resulting in an additional layer of cross talk among key TFs. Such cross
talk may direct the development of neuronal heterogeneity now known to exist in the
ARC. In this regard, it is interesting to note that Isl1 directs fate determination of
“specific” cell types by forming cell type-specific complexes with other TFs (Cho
et al. 2014; Habener et al. 2005; Lee and Pfaff 2001; Mazzoni et al. 2013; Peng et al.
2005) (Fig. 4.4b). As Isl1 is required for the generation of several distinct types of
arcuate neurons (Lee et al. 2016), Isl1 may partner with other TFs in specifying cell-
fate of each arcuate neuronal subtype.
4 Gene Regulatory Programs in the Development of Hypothalamic Arcuate. . . 95
The lack of early fate markers has prohibited us from deciphering the full spectrum
of activities that each TF exerts in the production of arcuate neuronal populations.
Most fate markers that are currently being used in the field, such as GHRH for
GHRH-neurons and AgRP for AgRP-neurons, represent late fate markers that
identify already committed or differentiated neurons. In addition, markers currently
used as early fate markers, label multiple neuronal types. For example, POMC marks
early ARC cells but lacks specificity given POMC-progenitors produce POMC-
neurons, AgRP-neurons as well as Kiss1-neurons (Padilla et al. 2010; Sanz et al.
2015). Similarly, we recently found that NPY+ progenitors likely give rise to AgRP-
and GHRH-neurons (Lee et al. 2018). The multiple lineages from POMC+ and
NPY+ precursors highlight the need for additional early fate markers, which will
enable us to understand the exact roles of each TF in specifying different arcuate
neuronal cell types. Future investigation into the ontogeny of arcuate neurons, such
as the relationship among POMC+ and NPY+ precursors and their progenies,
coupled with our understanding of the gene regulatory programs for arcuate neuronal
development, will reveal how different types of arcuate neurons are produced in the
right ratio during development to perform highly interconnected functions at post-
natal stages.
The recent advances in single cell RNA-seq (scRNA-seq) technology (Fig. 4.5)
will greatly facilitate the efforts to identify additional early fate markers of develop-
ing arcuate neurons, as this technology provides the transcriptome of individual
developing arcuate neurons. Using this technology in the hypothalamus at various
embryonic stages, we can identify additional early fate markers for each cell type. In
addition, developmental trajectories of each arcuate neuronal population and when
common progenitors begin to diverge to different arcuate neuronal lineages during
embryogenesis may be unraveled. Single cell RNA-seq can also be performed for
control and KO/cKO models for each TF, and their comparative analyses will
Fig. 4.5 Schematics for scRNA-seq. Dissociated cells are highly diluted so that only a single cell
mixed with reagents is encapsulated into nanoliter-sized GEMs (gel bead in emulsion) together with
a bead containing a barcoded oligonucleotide. GEMs are subjected to reverse transcription, and the
resulting cDNA is isolated and constructed into a library, followed by sequencing and analyses
96 J. W. Lee et al.
identify target genes of each TF. Integration of these results with the above described
CHIP-seq results will identify target genes that are directly controlled by the TF,
ultimately allowing us to fully reconstruct the gene regulatory programs that direct
the development of individual arcuate neurons.
Cre lines that can delete genes of interests only in a specific population of developing
arcuate neurons will allow us to investigate the developmental action of each TF in a
specific arcuate neuronal type independent of its action in other arcuate and
non-arcuate neuronal types. Notably, while Nkx2.1-Cre transgenic line has been
widely used for studying roles of genes of interests in the developing ARC, it deletes
genes in most neuronal types in the developing ARC (Xu et al. 2008). Moreover,
Nkx2.1 is also expressed in the thyroid, lung, and other parts of the brain (Lazzaro
et al. 1991). This specificity issue also extends to other Cre lines that delete genes of
interests in developing arcuate neurons, such as Npy-Cre (Milstein et al. 2015) and
Pomc-Cre (Balthasar et al. 2004). The promoters of early fate markers to be identified
from the aforementioned single cell RNA-seq approaches may allow us to create
more specific Cre lines. However, many of these early markers may still be expressed
in more than a single cell type. To help resolve this issue, an additional layer of
specificity can be achieved using the split-Cre system, in which co-expression of
catalytically inactive N-terminal and C-terminal fragments of Cre reactivates Cre
activity (Casanova et al. 2003; Hirrlinger et al. 2009) (Fig. 4.6). In this system, for
instance, two different promoters from early fate marker genes that are expressed in
arcuate cell types A and B (promoter 1) and B, C, and D (promoter 2), respectively,
will drive the expression of reconstituted active Cre in the overlapping cell type B
(Fig. 4.5).
Demonstrating a postnatal homeostatic phenotype that correlates with a develop-
mental deficit in an arcuate neuronal subtype is an excellent way to validate cell
specification. Notably, different arcuate neurons are connected to each other as well
Fig. 4.6 Schematics for the split-Cre system. To obtain higher specificity, expression of the
catalytically inactive N-terminal and C-terminal fragments of Cre is driven by two less-specific
promoters, respectively. Only in cell types in which both promoters are active, N-terminal and
C-terminal fragments of Cre are coexpressed restoring Cre activity
4 Gene Regulatory Programs in the Development of Hypothalamic Arcuate. . . 97
as with neurons in other brain regions, forming complex neural circuitries that
mediate homeostatic processes. Therefore, less-specific Cre lines, such as the
Nkx2.1-Cre transgenic line (Xu et al. 2008), pose a challenge for deciphering such
complex circuitry. For instance, mice with conditional deletions of Isl1 or Dlx1/2 in
developing arcuate neurons display developmental deficits in arcuate neurons that
control energy homeostasis and linear growth. Postnatally, these mice show dwarf-
ism as well as reduced energy expenditure and late-onset obesity (Lee et al. 2016,
2018). However, the specificity issues of Nkx2.1-Cre line make it difficult to
conclude that the postnatal phenotypes observed with Isl1- or Dlx1/2-cKO mice
are mainly due to the developmental deficits in arcuate neurons (Lee et al. 2016,
2018). For instance, Isl1 may also play a role in the development of the thyroid gland
(Westerlund et al. 2008). Thus, potential hypothyroidism resulting from the lack of
Isl1 may be the major culprit or have also contributed to the metabolic phenotypes
observed in the Isl1-cKO mice (Lee et al. 2016). These experiments highlight the
necessity for the development of more specific Cre lines that will enable us to
successfully link developmental deficits to adult ARC-related phenotypes.
Acknowledgments This work was supported by grants from NIH/NIDDK (R01 DK064678 and
R01 DK103661 to J.W.L.), the Korea Health Technology R&D Project through the Korea Health
Industry Development Institute, funded by the Ministry of Health & Welfare, Republic of Korea
(HI17C0447 to S.L.), and National Research Foundation of Korea (NRF-2012M3A9D1054705 to
S.L.).
98 J. W. Lee et al.
Key References
(continued)
4 Gene Regulatory Programs in the Development of Hypothalamic Arcuate. . . 99
References
Acampora D, Postiglione MP, Avantaggiato V, Di Bonito M, Vaccarino FM, Michaud J, Simeone
A (1999) Progressive impairment of developing neuroendocrine cell lineages in the hypothala-
mus of mice lacking the Orthopedia gene. Genes Dev 13:2787–2800
Altman J, Bayer SA (1978) Development of the diencephalon in the rat. II. Correlation of the
embryonic development of the hypothalamus with the time of origin of its neurons. J Comp
Neurol 182:973–993
Andrews GL, Yun K, Rubenstein JL, Mastick GS (2003) Dlx transcription factors regulate
differentiation of dopaminergic neurons of the ventral thalamus. Mol Cell Neurosci 23:107–120
Anthwal N, Pelling M, Claxton S, Mellitzer G, Collin C, Kessaris N, Richardson WD, Gradwohl G,
Ang SL (2013) Conditional deletion of neurogenin-3 using Nkx2.1iCre results in a mouse
model for the central control of feeding, activity and obesity. Dis Model Mech 6:1133–1145
Aujla PK, Naratadam GT, Xu L, Raetzman LT (2013) Notch/Rbpjkappa signaling regulates
progenitor maintenance and differentiation of hypothalamic arcuate neurons. Development
140:3511–3521
Bailey TL, Boden M, Buske FA, Frith M, Grant CE, Clementi L, Ren J, Li WW, Noble WS (2009)
MEME SUITE: tools for motif discovery and searching. Nucleic Acids Res 37:W202–W208
Balthasar N, Coppari R, McMinn J, Liu SM, Lee CE, Tang V, Kenny CD, McGovern RA, Chua SC
Jr, Elmquist JK et al (2004) Leptin receptor signaling in POMC neurons is required for normal
body weight homeostasis. Neuron 42:983–991
Barski A, Zhao K (2009) Genomic location analysis by ChIP-Seq. J Cell Biochem 107:11–18
Begley CG, Lipkowitz S, Gobel V, Mahon KA, Bertness V, Green AR, Gough NM, Kirsch IR
(1992) Molecular characterization of NSCL, a gene encoding a helix-loop-helix protein
expressed in the developing nervous system. Proc Natl Acad Sci USA 89:38–42
Biebermann H, Kuhnen P, Kleinau G, Krude H (2012) The neuroendocrine circuitry controlled by
POMC, MSH, and AGRP. Handb Exp Pharmacol 209:47–75
Bluet-Pajot MT, Tolle V, Zizzari P, Robert C, Hammond C, Mitchell V, Beauvillain JC, Viollet C,
Epelbaum J, Kordon C (2001) Growth hormone secretagogues and hypothalamic networks.
Endocrine 14:1–8
Byerly MS, Blackshaw S (2009) Vertebrate retina and hypothalamus development. Wiley
Interdiscip Rev Syst Biol Med 1:380–389
Campbell JN, Macosko EZ, Fenselau H, Pers TH, Lyubetskaya A, Tenen D, Goldman M,
Verstegen AM, Resch JM, McCarroll SA et al (2017) A molecular census of arcuate hypothala-
mus and median eminence cell types. Nat Neurosci 20:484–496
Casanova E, Lemberger T, Fehsenfeld S, Mantamadiotis T, Schutz G (2003) Alpha complementa-
tion in the Cre recombinase enzyme. Genesis 37:25–29
Chehab FF (2014) 20 years of leptin: leptin and reproduction: past milestones, present undertakings,
and future endeavors. J Endocrinol 223:T37–T48
Cho HH, Cargnin F, Kim Y, Lee B, Kwon RJ, Nam H, Shen R, Barnes AP, Lee JW, Lee S et al
(2014) Isl1 directly controls a cholinergic neuronal identity in the developing forebrain and
spinal cord by forming cell type-specific complexes. PLoS Genet 10:e1004280
100 J. W. Lee et al.
Cobos I, Borello U, Rubenstein JL (2007) Dlx transcription factors promote migration through
repression of axon and dendrite growth. Neuron 54:873–888
Cohen P, Rosenfeld RG (1994) Physiologic and clinical relevance of the insulin-like growth factor
binding proteins. Curr Opin Pediatr 6:462–467
Considine RV, Sinha MK, Heiman ML, Kriauciunas A, Stephens TW, Nyce MR, Ohannesian JP,
Marco CC, McKee LJ, Bauer TL et al (1996) Serum immunoreactive-leptin concentrations in
normal-weight and obese humans. N Engl J Med 334:292–295
Cremona M, Colombo E, Andreazzoli M, Cossu G, Broccoli V (2004) Bsx, an evolutionary
conserved brain specific homeoboX gene expressed in the septum, epiphysis, mammillary
bodies and arcuate nucleus. Gene Expr Patterns 4:47–51
De Rubeis S, He X, Goldberg AP, Poultney CS, Samocha K, Cicek AE, Kou Y, Liu L, Fromer M,
Walker S et al (2014) Synaptic, transcriptional and chromatin genes disrupted in autism. Nature
515:209–215
Dieguez C, Casanueva FF (1995) Influence of metabolic substrates and obesity on growth hormone
secretion. Trends Endocrinol Metab 6:55–59
Eisenstat DD, Liu JK, Mione M, Zhong W, Yu G, Anderson SA, Ghattas I, Puelles L, Rubenstein
JL (1999) DLX-1, DLX-2, and DLX-5 expression define distinct stages of basal forebrain
differentiation. J Comp Neurol 414:217–237
Ezzat S, Mader R, Fischer S, Yu S, Ackerley C, Asa SL (2006) An essential role for the
hematopoietic transcription factor Ikaros in hypothalamic-pituitary-mediated somatic growth.
Proc Natl Acad Sci USA 103:2214–2219
Gobel V, Lipkowitz S, Kozak CA, Kirsch IR (1992) NSCL-2: a basic domain helix-loop-helix gene
expressed in early neurogenesis. Cell Growth Differ 3:143–148
Good DJ, Porter FD, Mahon KA, Parlow AF, Westphal H, Kirsch IR (1997) Hypogonadism and
obesity in mice with a targeted deletion of the Nhlh2 gene. Nat Genet 15:397–401
Grinspoon S, Gulick T, Askari H, Landt M, Lee K, Anderson E, Ma Z, Vignati L, Bowsher R,
Herzog D et al (1996) Serum leptin levels in women with anorexia nervosa. J Clin Endocrinol
Metab 81:3861–3863
Habener JF, Kemp DM, Thomas MK (2005) Minireview: transcriptional regulation in pancreatic
development. Endocrinology 146:1025–1034
Hill JW, Elmquist JK, Elias CF (2008) Hypothalamic pathways linking energy balance and
reproduction. Am J Phys Endocrinol Metab 294:E827–E832
Hirrlinger J, Scheller A, Hirrlinger PG, Kellert B, Tang W, Wehr MC, Goebbels S, Reichenbach A,
Sprengel R, Rossner MJ et al (2009) Split-cre complementation indicates coincident activity of
different genes in vivo. PLoS One 4:e4286
Hrabovszky E (2014) Neuroanatomy of the human hypothalamic kisspeptin system. Neuroendocri-
nology 99:33–48
Hu Y, Yoshida T, Georgopoulos K (2017) Transcriptional circuits in B cell transformation. Curr
Opin Hematol 24:345–352
Ifft JD (1972) An autoradiographic study of the time of final division of neurons in rat hypothalamic
nuclei. J Comp Neurol 144:193–204
Iossifov I, O’Roak BJ, Sanders SJ, Ronemus M, Krumm N, Levy D, Stessman HA, Witherspoon
KT, Vives L, Patterson KE et al (2014) The contribution of de novo coding mutations to autism
spectrum disorder. Nature 515:216–221
Ishii Y, Bouret SG (2012) Embryonic birthdate of hypothalamic leptin-activated neurons in mice.
Endocrinology 153:3657–3667
Jing E, Nillni EA, Sanchez VC, Stuart RC, Good DJ (2004) Deletion of the Nhlh2 transcription
factor decreases the levels of the anorexigenic peptides alpha melanocyte-stimulating hormone
and thyrotropin-releasing hormone and implicates prohormone convertases I and II in obesity.
Endocrinology 145:1503–1513
Jones AR, Overly CC, Sunkin SM (2009) The Allen Brain Atlas: 5 years and beyond. Nat Rev
Neurosci 10:821–828
4 Gene Regulatory Programs in the Development of Hypothalamic Arcuate. . . 101
Kimura S, Hara Y, Pineau T, Fernandez-Salguero P, Fox CH, Ward JM, Gonzalez FJ (1996) The
T/ebp null mouse: thyroid-specific enhancer-binding protein is essential for the organogenesis
of the thyroid, lung, ventral forebrain, and pituitary. Genes Dev 10:60–69
Kruger M, Ruschke K, Braun T (2004) NSCL-1 and NSCL-2 synergistically determine the fate of
GnRH-1 neurons and control necdin gene expression. EMBO J 23:4353–4364
Kuperman Y, Weiss M, Dine J, Staikin K, Golani O, Ramot A, Nahum T, Kuhne C, Shemesh Y,
Wurst W et al (2016) CRFR1 in AgRP neurons modulates sympathetic nervous system activity
to adapt to cold stress and fasting. Cell Metab 23:1185–1199
Lam DD, de Souza FS, Nasif S, Yamashita M, Lopez-Leal R, Otero-Corchon V, Meece K,
Sampath H, Mercer AJ, Wardlaw SL et al (2015) Partially redundant enhancers cooperatively
maintain mammalian pomc expression above a critical functional threshold. PLoS Genet 11:
e1004935
Lasky JL, Wu H (2005) Notch signaling, brain development, and human disease. Pediatr Res
57:104r–109r
Lazzaro D, Price M, de Felice M, Di Lauro R (1991) The transcription factor TTF-1 is expressed at
the onset of thyroid and lung morphogenesis and in restricted regions of the foetal brain.
Development 113:1093–1104
Lee SK, Pfaff SL (2001) Transcriptional networks regulating neuronal identity in the developing
spinal cord. Nat Neurosci 4(Suppl):1183–1191
Lee B, Lee S, Lee SK, Lee JW (2016) The LIM-homeobox transcription factor Isl1 plays crucial
roles in the development of multiple arcuate nucleus neurons. Development 143:3763–3773
Lee B, Kim J, An T, Kim S, Patel EM, Raber J, Lee S-K, Lee S, Lee JW (2018) Dlx1/2 and Otp
coordinate the production of hypothalamic GHRH- and AgRP-neurons. Nat Commun 9:2026
Lein ES, Hawrylycz MJ, Ao N, Ayres M, Bensinger A, Bernard A, Boe AF, Boguski MS,
Brockway KS, Byrnes EJ et al (2007) Genome-wide atlas of gene expression in the adult
mouse brain. Nature 445:168–176
Li H, Zeitler PS, Valerius MT, Small K, Potter SS (1996) Gsh-1, an orphan Hox gene, is required
for normal pituitary development. EMBO J 15:714–724
Lintas C, Persico AM (2018) Unraveling molecular pathways shared by Kabuki and Kabuki-like
syndromes. Clin Genet 94(3–4):283-295
Long JE, Garel S, Alvarez-Dolado M, Yoshikawa K, Osumi N, Alvarez-Buylla A, Rubenstein JL
(2007) Dlx-dependent and -independent regulation of olfactory bulb interneuron differentiation.
J Neurosci Off J Soc Neurosci 27:3230–3243
Lu F, Kar D, Gruenig N, Zhang ZW, Cousins N, Rodgers HM, Swindell EC, Jamrich M,
Schuurmans C, Mathers PH et al (2013) Rax is a selector gene for mediobasal hypothalamic
cell types. J Neurosci Off J Soc Neurosci 33:259–272
Machanick P, Bailey TL (2011) MEME-ChIP: motif analysis of large DNA datasets. Bioinformat-
ics 27:1696–1697
Mannervik M, Nibu Y, Zhang H, Levine M (1999) Transcriptional coregulators in development.
Science 284:606–609
Marin O, Anderson SA, Rubenstein JL (2000) Origin and molecular specification of striatal
interneurons. J Neurosci Off J Soc Neurosci 20:6063–6076
Marin O, Baker J, Puelles L, Rubenstein JL (2002) Patterning of the basal telencephalon and
hypothalamus is essential for guidance of cortical projections. Development (Cambridge,
England) 129:761–773
Martinez-Ferre A, Martinez S (2012) Molecular regionalization of the diencephalon. Front
Neurosci 6:73
Mazzoni EO, Mahony S, Closser M, Morrison CA, Nedelec S, Williams DJ, An D, Gifford DK,
Wichterle H (2013) Synergistic binding of transcription factors to cell-specific enhancers
programs motor neuron identity. Nat Neurosci 16:1219–1227
McNay DE, Pelling M, Claxton S, Guillemot F, Ang SL (2006) Mash1 is required for generic and
subtype differentiation of hypothalamic neuroendocrine cells. Mol Endocrinol 20:1623–1632
102 J. W. Lee et al.
Michaelson JJ, Shi Y, Gujral M, Zheng H, Malhotra D, Jin X, Jian M, Liu G, Greer D, Bhandari A
et al (2012) Whole-genome sequencing in autism identifies hot spots for de novo germline
mutation. Cell 151:1431–1442
Milstein AD, Bloss EB, Apostolides PF, Vaidya SP, Dilly GA, Zemelman BV, Magee JC (2015)
Inhibitory gating of input comparison in the CA1 microcircuit. Neuron 87:1274–1289
Muranishi Y, Terada K, Furukawa T (2012) An essential role for Rax in retina and neuroendocrine
system development. Develop Growth Differ 54:341–348
Nasif S, de Souza FS, Gonzalez LE, Yamashita M, Orquera DP, Low MJ, Rubinstein M (2015) Islet
1 specifies the identity of hypothalamic melanocortin neurons and is critical for normal food
intake and adiposity in adulthood. Proc Natl Acad Sci USA 112:E1861–E1870
Niikawa N, Matsuura N, Fukushima Y, Ohsawa T, Kajii T (1981) Kabuki make-up syndrome: a
syndrome of mental retardation, unusual facies, large and protruding ears, and postnatal growth
deficiency. J Pediatr 99:565–569
O’Roak BJ, Vives L, Girirajan S, Karakoc E, Krumm N, Coe BP, Levy R, Ko A, Lee C, Smith JD
et al (2012) Sporadic autism exomes reveal a highly interconnected protein network of de novo
mutations. Nature 485:246–250
Padilla SL, Carmody JS, Zeltser LM (2010) Pomc-expressing progenitors give rise to antagonistic
neuronal populations in hypothalamic feeding circuits. Nat Med 16:403–405
Pan Q, Li C, Xiao J, Kimura S, Rubenstein J, Puelles L, Minoo P (2004) In vivo characterization of
the Nkx2.1 promoter/enhancer elements in transgenic mice. Gene 331:73–82
Pelling M, Anthwal N, McNay D, Gradwohl G, Leiter AB, Guillemot F, Ang SL (2011) Differential
requirements for neurogenin 3 in the development of POMC and NPY neurons in the hypothal-
amus. Dev Biol 349:406–416
Peng SY, Wang WP, Meng J, Li T, Zhang H, Li YM, Chen P, Ma KT, Zhou CY (2005) ISL1
physically interacts with BETA2 to promote insulin gene transcriptional synergy in non-beta
cells. Biochim Biophys Acta 1731:154–159
Petryniak MA, Potter GB, Rowitch DH, Rubenstein JL (2007) Dlx1 and Dlx2 control neuronal
versus oligodendroglial cell fate acquisition in the developing forebrain. Neuron 55:417–433
Price M, Lazzaro D, Pohl T, Mattei MG, Ruther U, Olivo JC, Duboule D, Di Lauro R (1992)
Regional expression of the homeobox gene Nkx-2.2 in the developing mammalian forebrain.
Neuron 8:241–255
Qiu M, Bulfone A, Martinez S, Meneses JJ, Shimamura K, Pedersen RA, Rubenstein JL (1995)
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
9:2523–2538
Rowitch DH, Kriegstein AR (2010) Developmental genetics of vertebrate glial-cell specification.
Nature 468:214–222
Saino-Saito S, Berlin R, Baker H (2003) Dlx-1 and Dlx-2 expression in the adult mouse brain:
relationship to dopaminergic phenotypic regulation. J Comp Neurol 461:18–30
Sakkou M, Wiedmer P, Anlag K, Hamm A, Seuntjens E, Ettwiller L, Tschop MH, Treier M (2007)
A role for brain-specific homeobox factor Bsx in the control of hyperphagia and locomotory
behavior. Cell Metab 5:450–463
Sanz E, Quintana A, Deem JD, Steiner RA, Palmiter RD, McKnight GS (2015) Fertility-regulating
Kiss1 neurons arise from hypothalamic POMC-expressing progenitors. J Neurosci Off J Soc
Neurosci 35:5549–5556
Schmid T, Gunther S, Mendler L, Braun T (2013) Loss of NSCL-2 in gonadotropin releasing
hormone neurons leads to reduction of pro-opiomelanocortin neurons in specific hypothalamic
nuclei and causes visceral obesity. J Neurosci Off J Soc Neurosci 33:10459–10470
Shimogori T, Lee DA, Miranda-Angulo A, Yang Y, Wang H, Jiang L, Yoshida AC, Kataoka A,
Mashiko H, Avetisyan M et al (2010) A genomic atlas of mouse hypothalamic development.
Nat Neurosci 13:767–775
4 Gene Regulatory Programs in the Development of Hypothalamic Arcuate. . . 103
Abstract
The daily cycle of sleep and wake governs all aspects of behavior and is under tight
homeostatic regulation. The core circadian oscillator in turn restricts sleep to
specific intervals of the solar day. The neurons that regulate both sleep and
circadian timing reside largely in the hypothalamus, and are linked in a complex,
reciprocally connected, and still incompletely characterized network. While our
knowledge of the neural circuitry controlling circadian rhythms and sleep has
advanced considerably, our understanding of the mechanisms by which they are
assembled during embryonic and postnatal development lag far behind. In this
chapter, we review advances in the understanding of hypothalamic circuitry
controlling circadian rhythms and sleep and our current knowledge of the gene
regulatory networks that guide their development. In addition, we discuss the
D. W. T. Kim
Solomon H. Snyder Department of Neuroscience, Johns Hopkins University School of Medicine,
Baltimore, MD, USA
S. Blackshaw (*)
Solomon H. Snyder Department of Neuroscience, Johns Hopkins University School of Medicine,
Baltimore, MD, USA
Department of Ophthalmology, Johns Hopkins University School of Medicine, Baltimore, MD,
USA
Department of Neurology, Johns Hopkins University School of Medicine, Baltimore, MD, USA
Center for Human Systems Biology, Johns Hopkins University School of Medicine, Baltimore,
MD, USA
Institute for Cell Engineering, Johns Hopkins University School of Medicine, Baltimore, MD, USA
e-mail: [email protected]
Keywords
Circadian · Development · Hypocretin · Sleep · Suprachiasmatic · Timing ·
Transcription factor · Zona incerta
5.1.1 Introduction
The hypothalamus has long been known to be a central and critical regulator of
sleep. Focal lesions to the anterior hypothalamus and preoptic area often result in
insomnia, while lesions to the posterior hypothalamus result in lethargy and
hypersomnia (Scammell et al. 2017; Weber and Dan 2016). Several distinct neural
processes interact to regulate sleep. First of these is the neural circuitry that directly
regulates sleep/wake transitions. Sleep and wakefulness are mutually exclusive
states, and transitions between these states occur rapidly, with “half asleep” being
a state that lasts only seconds. Reciprocal transitions between wake and sleep can
occur frequently, particularly in species such as mice and human infants. An
analogous process in turn mediates rapid, biphasic, and bidirectional transitions
between non-rapid eye movement (NREM) and rapid eye movement (REM) sleep
phases. The neural circuitry that controls sleep and wake involves three major
components. The first and most extensively studied directly regulates the transition
between wake, NREM, and REM sleep. The second and third, which are much less
well understood, regulate homeostatic sleep pressure, and circadian changes in sleep
pressure, respectively. Sensory signals and changes in internal states can in turn
modulate sleep/wake transitions. An overview of hypothalamic cell types that
control sleep, wakefulness, and biological timing is shown in Fig. 5.1.
Recent studies, particularly since the advent of optogenetics, have identified several
hypothalamic neuronal subtypes whose activity is sleep regulated, and which rapidly
induce either wakefulness or sleep following activation or inhibition (Fig. 5.2a).
Primary wake-promoting neurons act by sending widespread, and typically excit-
atory, projections throughout the forebrain. In the hypothalamus, these include the
histaminergic neurons of the tuberomammillary nucleus (TMN), whose wake-
promoting signals are blunted by conventional antihistamines; orexinergic neurons
of the lateral hypothalamus (LH), which degenerate in narcoleptic patients; and other
populations such as glutamatergic neurons of the supramammilary nucleus (Pedersen
5 Winding the Clock: Development of Hypothalamic Structures Controlling. . . 107
Fig. 5.1 Locations of hypothalamic sleep and wake-promoting neurons. (a) Sleep-promoting
neurons of the preoptic area. (b) Sleep and wake-promoting neurons of the dorsolateral hypothala-
mus. (c) Neuronal subtypes of the suprachiasmatic nucleus regulating biological timing. (d) Wake-
promoting neurons of the posteroventral hypothalamus. Diagrams of transverse sections through
hyopothalamus, centered on third ventricle (black). Abbreviations: Avp arginine vasopressin, Cck
cholecystokinin, Crh corticotropin-releasing hormone, DMH dorsomedial hypothalamic nucleus,
GABA gamma-amino butyric acid, Gal galanin, Grp gastrin-related peptide, Hdc histamine decar-
boxylase, LH lateral hypothalamus, Pmch pro-melanin concentrating hormone, SCN
suprachiasmatic nucleus, Tac1 tachykinin 1, TMN tuberomammillary nucleus, Vip vasoactive
intestinal peptide, VLPO ventrolateral preoptic area, ZIv zona incerta, ventral region
Fig. 5.2 Connections and actions of hypothalamic sleep-regulating neurons. (a) Schematic of
hypothalamic neurons directly regulating sleep and wake. (b) Pre- and postsynaptic connections of
sleep pressure-regulating Lhx6-expressing neurons of the zona incerta. (c) Delayed effects on sleep
of chemogenetic-dependent activation and inhibition of Cre-dependent DREADD constructs deliv-
ered Lhx6-expressing neurons of the zona incerta by bilateral stereotaxic injection [adapted from
Liu et al. (2017)]. (a, b) Sagittal brain sections; (c) coronal brain section. Abbreviations: CNO
clozapine-N-oxide, DB ventral diagonal band, DMH dorsomedial hypothalamic nucleus, DREADD
designer receptor exclusively activated by designer drugs, DRV dorsal raphe nucleus, ventral
region, LH lateral hypothalamus, LS lateral septum, POA preoptic area, PPTG pedunculopontine
tegmental nucleus, REM rapid eye movement, RtTg rostral tegmental nucleus, SNPC substantia
nigra, pars compacta, TMN tuberomammillary nucleus, vlPAG ventrolateral periaqueductal grey,
VTA ventral tegmental area, ZIv zona incerta, ventral region, ZT zeitgeber time
Fig. 5.3 Suprachiasmatic nucleus (SCN): circadian rhythms and neuronal differentiation. (a)
Overview of light-dependent regulation of the core circadian oscillator in the SCN. Light is relayed
to the SCN via intrinsically photosensitive retinal ganglion cells to regulate rhythmic gene expres-
sion. SCN neurons. (b) Transcriptional regulatory networks (left) controlling SCN differentiation,
(right) and controlling expression of clock-regulating neuropeptides (Avp arginine vasopressin, Grp
gastrin releasing peptide, Vip vasoactive intestinal peptide)
Like all cells in the adult, a biochemical circadian oscillator is active in neurons of
the SCN (Fig. 5.3). This takes the form of a negative feedback loop, with a
heterodimeric transcriptional complex consisting of the transcription factors Bmal1
and Clock, which drive the expression of Per and Cry genes, which in turn feedback
to inhibit their own transcription. This core oscillator is reinforced and stabilized by
linked transcriptional negative feedback loops, involving the nuclear hormone
receptor superfamily members Rora, Rorb, and Nr1d1 (Partch et al. 2014). In
sharp contrast to virtually all other tissues, however, the cellular clocks of SCN
neurons undergo spontaneous synchronization when grown in dispersed cell culture
and remain robustly synchronized when maintained ex vivo in intact organ culture
(Mohawk and Takahashi 2011). Moreover, while cellular clocks in other tissues are
readily reset and synchronized by a range of external signals—including
5 Winding the Clock: Development of Hypothalamic Structures Controlling. . . 111
temperature, metabolites, and growth factors—cells in the SCN robustly retain their
phase and synchrony under a broad range of treatments. These unique features arise
at least in part as a result of the organization of local neural circuitry in the SCN. Like
the ZI, the SCN consists of interconnected GABAergic neurons, and this extensive
collateral inhibition buffers the SCN against perturbation. Inhibition of neuronal
excitability in the SCN reverses this stability, allowing cellular oscillators to be
readily reset by light and temperature signals that would normally produce no
change in phase (Buhr et al. 2010; Yamaguchi et al. 2013). SCN neurons also
release a diverse array of neuropeptides—including Vip, Avp, Grp, Penk, and
Vgf—which likewise contribute to phase synchrony and stabilization. While the
molecular mechanisms and, to a lesser extent, the local circuitry that controls
circadian timekeeping in the SCN, are now fairly well understood, the neural
circuitry that relays these signals from the SCN to other brain regions is not.
Finally, a range of sensory and internal signals can modulate sleep, usually by
directly promoting wakefulness without altering either homeostatic or circadian
sleep pressure. Sensory signals include loud noises, temperature extremes, and
pain, with the most complex of these in its effects being light. Light directly inhibits
(in the case of diurnal mammals) or promotes (in the case of nocturnal mammals) the
onset of sleep. Moreover, while circadian sleep pressure is highly robust in the face
of most sensory and internal signals, it can be reset by light signals relayed directly
from intrinsically photosensitive retinal ganglion cells (ipRGCs) (Hattar et al. 2002).
These ipRGCs directly activate neurons in the ventral SCN, which in turn leads to a
gradual and progressive resetting of cellular SCN clocks. This phase shift occurs at
the rate of approximately 1 h per day, with misalignment of circadian and clock time
leading to the phenomenon of jet lag (Sack et al. 2007). Some examples of internal
states that acutely modulate sleep (though not sleep pressure) include sympathetic
and behavior arousal, such as hunger and thirst, or psychotropic drugs that directly
target arousal circuitry (Saper et al. 2010). Despite the strong and rapid effects of
these signals on sleep, surprisingly little is known about the circuitry mediating
clock-dependent regulation of sleep. In addition, virtually nothing is known about
the mechanisms by which changes in sleep state in turn induce sleep-dependent
changes in internal states and sensory processing.
rodents between postnatal day 12 and 20 (Daszuta and Gambarelli 1985; Frank and
Heller 1997; Nelson et al. 2013). However, analysis of other sleep-regulated
behaviors, such as muscle tone measured by electromyogram (EMG), suggest that
sleep/wake transitions may be evident in rodents as early as postnatal day 2 (Gall
et al. 2008). Light-dependent regulation of sleep–wake cycles is evident by weaning
in rodents (Nelson et al. 2013; Sack et al. 2007) and begin to emerge by 10–12 weeks
postnatal in humans (Davis et al. 2004). Circadian control of circulating cortisol and
core body temperature also becomes evident around this age (Joseph et al. 2015),
and this is likely the point at which sleep/wake cycles also come under circadian
regulation (Blumberg et al. 2014; Galland et al. 2012). This process occurs long after
molecular rhythms are first detected in the SCN. These are first observed by E15,
coincident with expression of the first SCN-enriched neuropeptides, and become
strong and well consolidated by postnatal day 2 (Carmona-Alcocer et al. 2018),
implying a substantial delay in the development of efferent neural circuitry relaying
SCN output. There is a gradual shift in sleep homeostasis through life, with progres-
sively less time spent in sleep, with major reductions observed first in REM, then in
deep NREM (Dzierzewski et al. 2018; Miles and Dement 1980). Sleep/wake phases
are relatively phase-advanced in childhood, phase-delayed in adolescence and early
adulthood, and phase-advanced and less well consolidated in the elderly. Both
Mendelian and multifactorial genetic variation control the extent to which sleep/
wake cycles are phase-advanced or phase-delayed relative to the solar day, giving
rise to what are known as chronotypes (Roenneberg et al. 2003). This chronotypic
variation in sleep/wake timing peaks in early adulthood and declines slowly through-
out the lifespan (Fischer et al. 2017).
5.2.1 Introduction
5.2.2.1 Shh
Shh (sonic hedgehog) has been studied more extensively than any other extrinsic
factor in the context of hypothalamic development. Shh expression in the notochord,
and its anterior derivative the prechordal plate, leads to induction of Shh expression
in the ventral floor plate all along the neuraxis, including in hypothalamic
neuroepithelium. Loss of function of Shh at this stage blocks development of all
hypothalamic structures (Blaess et al. 2014). Soon thereafter, Shh expression is
induced in a more dorsally located region known at the basal plate domain, and
floorplate expression is suppressed. Selective loss of function of Shh in the basal
plate domain prior to the onset of neurogenesis leads to the loss of all anterior and
tuberal hypothalamic markers, along with severe defects in neural progenitor prolif-
eration. This prevents development of all sleep-associated neurons, with the excep-
tion of supramammilary nucleus neurons (Shimogori et al. 2010).
Fig. 5.4 Molecular organization of the hypothalamus and prethalamus in the embryonic (day E12)
mouse brain. (a) Parasagittal schematic (adapted from Shimogori et al. (2010)). (b) Selected
molecular markers for the indicated regions. Abbreviations: from top to bottom, left group: ZLI
zona limitans intrathalamica, PreThal prethalamus, EmThal thalamic eminence, PVN hypothalamic
paraventricular nucleus, vAH ventral anterior hypothalamus, ID intrahypothalamic diagonal, TT
tuberomammillary terminal, VMH ventromedial hypothalamus, ArcN arcuate nucleus, PMN
premammillary nucleus, MM mammillary nucleus, SMM supramammillary nucleus, Shh sonic
hedgehog
Despite its small size (~20,000 neurons total), the physiological importance and
compact, easily identifiable structure of the SCN has resulted in the transcription
factors that control its development being better characterized than any other struc-
ture in the hypothalamus. Nearly a decade of work has made considerable progress
in identifying a hierarchical network of transcription factors that controls specifica-
tion, differentiation, and adult function of the SCN (Fig. 5.4).
(EmThal) (Roy et al. 2013). In the neuroretina, the repressive function of Lhx2
persists even after neurogenesis has begun (de Melo et al. 2016; Roy et al. 2013).
Table 5.1 Behavioral defects in mouse mutants that disrupt development of hypothalamic neurons
controlling sleep and circadian timing
Mutant line Cellular phenotype Behavioral phenotype References
Six3-Cre: Broadly decreased Loss of SCN-enriched Bedont et al.
Lhx1lox/lox expression of neuropeptides. Disruption of (2014, 2017)
SCN-enriched activity, sleep, and body
neuropeptides. temperature rhythms. Loss of
resistance to temperature-
dependent resetting of SCN clock.
Loss of light-dependent regulation
of sleep.
Rora-Cre: Decreased expression Moderate defects in circadian Hatori et al.
Lhx1lox/lox of a subset of activity rhythms. Enhanced light- (2014)
SCN-enriched dependent resetting of circadian
neuropeptides. phase.
Zfhx3sci/sci Broadly reduced Defective circadian activity Balzani et al.
expression of rhythms. Altered sleep and interval (2016) and
SCN-enriched timing. Parsons et al.
neuropeptides. (2015)
Ubc- Not determined. Disrupted circadian activity Wilcox et al.
CreER; rhythms. (2017)
Zfhx3lox/lox
Foxd1- Loss of hypothalamic Reduced sleep time, increased Liu et al. (2017)
Cre; Lhx6+ neurons. wake time.
Lhx6lox/lox
Lhx9 / Reduction in Hcrt+ Reduced locomotor activity and Dalal et al.
neuron number. increased wake time. (2013)
function in postmitotic SCN neural precursors results in a milder but broadly similar
phenotype (Hatori et al. 2014). Lhx1 expression is maintained in adult SCN
(VanDunk et al. 2011) and is regulated by both the circadian clock and by light
(Hatori et al. 2014).
Changes in gene expression in Lhx1-deficient mice appear to reflect both the
direct and indirect effects of Lhx1. Lhx1 directly regulates Vip expression (Hatori
et al. 2014) and Vip shows the strongest effects of all SCN-enriched neuropeptides
on regulation of cellular and behavioral rhythms (Maywood et al. 2011). Loss of
function of Vip leads to defects in circadian rhythmicity, and loss of expression of a
large fraction of Lhx1-dependent genes, including neuropeptides such as Avp and
Penk, implying that Lhx1-dependent regulation of Vip expression accounts for these
effects (Bedont et al. 2017). However, expression of other SCN-enriched
neuropeptides such as Grp, Nms, and Prok2 is unaffected by Vip loss of function,
implying that Lhx1 may act as a master regulator of SCN neuropeptide expression
(Table 5.1).
Much less is known about transcription factors that control development of sleep-
regulating neurons in the posterior hypothalamus, with work on this topic having
mostly focused on the function of the LIM homeodomain factors Lhx6 and Lhx9.
Lhx6 and Lhx9 are expressed in adjacent, nonoverlapping regions of the posterolat-
eral hypothalamus from E11 into adulthood in mouse (Shimogori et al. 2010). Lhx6
expression is ultimately restricted to GABAergic neurons in the ventral ZI, DMH,
and LH (Liu et al. 2017), while Lhx9 is restricted to excitatory orexinergic neurons
(Dalal et al. 2013). Although loss of function of Lhx6 disrupts the function of sleep-
promoting ZI neurons, exactly how this gene regulates the specification and/or
function of these cells is still unclear. Lhx9 expression is detected in orexinergic
cells long before orexin itself is expressed (Shimogori et al. 2010), and in zebra fish,
Lhx9 directly regulates orexin expression (Liu et al. 2017). In mice, loss of function
of Lhx9 leads to a reduction in both the levels of orexin expression and a 30%
reduction in the number of orexigenic neurons (Dalal et al. 2013) (Table 5.1). Lhx9-
deficient zebra fish shows a complete loss of orexin expression (Liu et al. 2017), In
contrast, overexpression of Lhx9 in both species increases the number of orexinergic
neurons, but in mice did so only in posterior hypothalamus (Liu et al. 2015).
Moreover, viral overexpression of Lhx9 in adult Lhx9-deficient mice does not rescue
118 D. W. T. Kim and S. Blackshaw
defects in hypocretin expression (Liu et al. 2015). Gene expression profiling analysis
of orexinergic neurons has identified other transcription factors that are selectively
expressed in orexinergic neurons, although their role in differentiation and function
of these cells has not been investigated.
Mice mutant for Lhx1 have provided important insights into the molecular
mechanisms regulating circadian synchrony and behavioral output of the SCN.
SCN neurons express a diverse collection of neuropeptides, including Vip, Avp,
Grp, Prok2, Nms, Penk, and Vgf (Abrahamson and Moore 2001). Expression of
many of these is regulated by both circadian time and light, and they have been
proposed to both regulate the synchrony of cellular oscillators within the SCN and
mediate output from the SCN to other brain regions, allowing them to keep time with
5 Winding the Clock: Development of Hypothalamic Structures Controlling. . . 119
the master clock (Aton and Herzog 2005). Genetic and pharmacological analysis
have identified a hierarchy of function among these neuropeptides, at least with
respect to synchronization of cellular oscillators, with Vip being dominant over Avp,
and Avp over Grp (Maywood et al. 2011). The sheer number of candidate
neuropeptides, however, has not made it feasible to investigate the role of
neuropeptides in central clock function more generally, as opposed to investigating
the function of individual neuropeptides.
Disruption of Lhx1 function in hypothalamic neuroepithelial cells leads to a near-
complete loss of expression of all SCN-enriched neuropeptides without grossly
affecting either regional organization or GABAergic neurotransmission in the
SCN (Bedont et al. 2014). Cellular rhythms within the SCN remain locally
synchronized in explants from Lhx1-deficient mice, but desynchrony is observed
among different regions within the SCN, and this becomes more pronounced over
time (Bedont et al. 2014). This is reflected in the behavioral phenotype of mutant
mice, which though grossly normal, show either phase splitting or lack of circadian
rhythmicity in locomotor and body temperature rhythms (Bedont et al. 2014).
Interestingly, the loss of neuropeptide expression in Lhx1-deficient mice renders
the core behavioral oscillator highly susceptible to perturbation and resetting. SCN
explants from Lhx1-deficient mice lose their resistance to resetting by temperature
pulses, essentially phenocopying samples in which neuronal excitability is
suppressed using tetrodotoxin (Bedont et al. 2017). Moreover, induction of fever
by rhythmic injections of lipopolysaccharide is sufficient to reset both core body
temperature and locomotor rhythms of these mutant mice (Bedont et al. 2017). These
same animals are also rendered susceptible to phase shifts induced by
intracerebroventricular canulation of SCN-enriched neuropeptides that do not nor-
mally influence behavioral rhythms (Bedont et al. 2014). Mice showing later loss of
function of Lhx1 show milder phenotypes that also reflect reduced coupling of
cellular oscillators in the SCN (Hatori et al. 2014). Taken together, these data
show that Lhx1-regulated neuropeptide gene expression plays a central role in the
stabilization of cellular oscillators within the SCN, buffering them against resetting
by internal signals and also in mediating efferent signaling from the SCN to other
brain regions.
1 h (Jud et al. 2005). This implies that SCN-enriched neuropeptides are essential for
relaying light signals from ipRGCs to cellular clocks in the SCN. These phenotypes
are not observed when Lhx1 is selectively deleted in SCN neural precursors, with
these mutants showing enhanced resetting of circadian activity rhythms by light
pulse (Hatori et al. 2014).
Early deletion of Lhx1 leads to an unexpected loss of acute light-dependent
regulation of sleep (Bedont et al. 2017). This phenotype is distinct from light-
dependent regulation of the core circadian clock, which is completely disrupted in
these mutants, and identifies an unexpected role for the SCN in direct regulation of
sleep/wake transitions. These mutants do not show any changes in sleep homeosta-
sis. The neural circuitry downstream of the SCN that controls acute light-dependent
regulation of sleep remains unclear, however.
Like Lhx1, Zfhx3 has an essential role in control of circadian locomotor rhythms and
sleep. Dominant ENU-induced missense mutations in Zfhx3 show shortened bio-
chemical and circadian activity periods, while these were lengthened by Zfhx3
knockdown (Parsons et al. 2015). In contrast, selective knockout of Zhfx3 in adult
animals results in either shortened rhythms or else led to complete arrhythmicity.
Animals that retained rhythmicity rapidly entrained to phase-advanced (but not
phase-delayed) light cycles, implying that coupling among cellular SCN oscillators
was attenuated (Parsons et al. 2015).
Missense mutations in Zfhx3 mutants also showed evidence for shortened interval
timing, when trained to respond with a 10s delay to a conditioned stimulus (Balzani
et al. 2016), suggesting that they may show a more general defect in biological
timing. Zfhx3 missense mutants also show a reduction in rebound NREM sleep
following 6 h. of sleep deprivation but do not show any significant differences in
sleep homeostasis under baseline conditions (Balzani et al. 2016). Light-dependent
regulation of sleep is also intact in these mutants (Balzani et al. 2016). Since these
data were obtained using constitutive mutants, and Zfhx3 is expressed broadly in the
forebrain, it is unclear whether defects in interval timing and sleep homeostasis are a
direct result of loss of Zfhx3 in the SCN or other hypothalamic structures.
Lhx9-deficient mice show reduced locomotor activity and increased rest time,
although sleep was not directly investigated (Dalal et al. 2013). Unfortunately, in
zebra fish, the effects of gain and loss of function of Lhx9 on sleep behavior have
also not been investigated (Liu et al. 2015)—but since Lhx9 is crucial for Hcrt-
expressing neuronal differentiation, we expect there is a defect in sleep behavior.
It is unclear how evolutionarily conserved this circuitry is likely to be. The location
of the central circadian clock in nonmammalian vertebrates is still unclear, with the
location and function of SCN being ambiguous even in birds. Some conservation of
key sleep-regulatory neurons (e.g., Hcrt, Lhx6, Pmch) is observed in zebra fish (Liu
et al. 2015; Wolf and Ryu 2013; Xie et al. 2017). In some cases, these have also been
shown to be functionally homologous (Liu et al. 2015). Studies of nonmammalian
vertebrates will help resolve this question—in particular, a systematic and unbiased
analysis of gene expression in developing hypothalamus like that performed in
developing mouse hypothalamus (Shimogori et al. 2010).
Recent years have seen rapid improvement in protocols for directed differentiation of
human ES and iPS cells into a broad range of neuronal subtypes, including a number
of mediobasal hypothalamic cell types (Merkle et al. 2015; Ogawa et al. 2018; Wang
et al. 2015; Wataya et al. 2008; Yao et al. 2017). The more recent development of
hypothalamic organoid preparations (Qian et al. 2016, 2018) raises the possibility of
potentially analyzing the formation and function of sleep-regulating structures such
as the SCN or ZI in vitro.
Before this can come about, however, we will need to know much more about the
basic molecular mechanisms controlling hypothalamic patterning and neurogenesis.
In particular, we lack a clear understanding of general principles guiding organiza-
tion and function of even the SCN. For instance, we have no idea how individual
transcription factors that have been found to be necessary to regulate rhythmicity
actually do so. What are their genomic targets? Is expression of these factors
sufficient for initiation of rhythmicity and synchrony of cellular clocks, as well as
necessary? The ability to perform gain and loss of function analysis in organoid
preparations (Eldred et al. 2018; Mariani et al. 2015), in combination with
techniques for globally profiling transcription factor target sites such as ATAC-
5 Winding the Clock: Development of Hypothalamic Structures Controlling. . . 123
Key References
References
Abrahamson EE, Moore RY (2001) Suprachiasmatic nucleus in the mouse: retinal innervation,
intrinsic organization and efferent projections. Brain Res 916:172–191
Aton SJ, Herzog ED (2005) Come together, right. . .now: synchronization of rhythms in a mamma-
lian circadian clock. Neuron 48:531–534
Aujla PK, Naratadam GT, Xu L, Raetzman LT (2013) Notch/Rbpjkappa signaling regulates
progenitor maintenance and differentiation of hypothalamic arcuate neurons. Development
140:3511–3521
Balzani E, Lassi G, Maggi S, Sethi S, Parsons MJ, Simon M, Nolan PM, Tucci V (2016) The Zfhx3-
mediated axis regulates sleep and interval timing in mice. Cell Rep 16:615–621
Bedont JL, LeGates TA, Slat EA, Byerly MS, Wang H, Hu J, Rupp AC, Qian J, Wong GW, Herzog
ED et al (2014) Lhx1 controls terminal differentiation and circadian function of the
suprachiasmatic nucleus. Cell Rep 7:609–622
Bedont JL, LeGates TA, Buhr E, Bathini A, Ling JP, Bell B, Wu MN, Wong PC, Van Gelder RN,
Mongrain V et al (2017) An LHX1-regulated transcriptional network controls sleep/wake
coupling and thermal resistance of the central circadian clockworks. Curr Biol 27:128–136
Biehl MJ, Kaylan KB, Thompson RJ, Gonzalez RV, Weis KE, Underhill GH, Raetzman LT (2018)
Cellular fate decisions in the developing female anteroventral periventricular nucleus are
regulated by canonical Notch signaling. Dev Biol 442:87–100
Blaess S, Szabo N, Haddad-Tovolli R, Zhou X, Alvarez-Bolado G (2014) Sonic hedgehog
signaling in the development of the mouse hypothalamus. Front Neuroanat 8:156
Blumberg MS, Gall AJ, Todd WD (2014) The development of sleep-wake rhythms and the search
for elemental circuits in the infant brain. Behav Neurosci 128:250–263
Borbely AA (1982) A two process model of sleep regulation. Hum Neurobiol 1:195–204
124 D. W. T. Kim and S. Blackshaw
Brunner DP, Dijk DJ, Tobler I, Borbely AA (1990) Effect of partial sleep deprivation on sleep
stages and EEG power spectra: evidence for non-REM and REM sleep homeostasis.
Electroencephalogr Clin Neurophysiol 75:492–499
Buhr ED, Yoo SH, Takahashi JS (2010) Temperature as a universal resetting cue for mammalian
circadian oscillators. Science 330:379–385
Campbell JN, Macosko EZ, Fenselau H, Pers TH, Lyubetskaya A, Tenen D, Goldman M,
Verstegen AM, Resch JM, McCarroll SA et al (2017) A molecular census of arcuate hypothala-
mus and median eminence cell types. Nat Neurosci 20:484–496
Carmona-Alcocer V, Abel JH, Sun TC, Petzold LR, Doyle FJ 3rd, Simms CL, Herzog ED (2018)
Ontogeny of circadian rhythms and synchrony in the suprachiasmatic nucleus. J Neurosci
38:1326–1334
Chen R, Wu X, Jiang L, Zhang Y (2017) Single-cell RNA-Seq reveals hypothalamic cell diversity.
Cell Rep 18:3227–3241
Chen KS, Xu M, Zhang Z, Chang WC, Gaj T, Schaffer DV, Dan Y (2018) A hypothalamic switch
for REM and non-REM sleep. Neuron 97(1168–1176):e1164
Chung S, Weber F, Zhong P, Tan CL, Nguyen TN, Beier KT, Hormann N, Chang WC, Zhang Z,
Do JP et al (2017) Identification of preoptic sleep neurons using retrograde labelling and gene
profiling. Nature 545:477–481
Clark DD, Gorman MR, Hatori M, Meadows JD, Panda S, Mellon PL (2013) Aberrant development
of the suprachiasmatic nucleus and circadian rhythms in mice lacking the homeodomain protein
Six6. J Biol Rhythm 28:15–25
Daan S, Beersma DG, Borbely AA (1984) Timing of human sleep: recovery process gated by a
circadian pacemaker. Am J Phys 246:R161–R183
Dalal J, Roh JH, Maloney SE, Akuffo A, Shah S, Yuan H, Wamsley B, Jones WB, de Guzman
Strong C, Gray PA et al (2013) Translational profiling of hypocretin neurons identifies candidate
molecules for sleep regulation. Genes Dev 27:565–578
Daszuta A, Gambarelli F (1985) Early postnatal development of EEG and sleep-waking cycle in
two inbred mouse strains. Brain Res 354:39–47
Davis KF, Parker KP, Montgomery GL (2004) Sleep in infants and young children: part one:
normal sleep. J Pediatr Health Care 18:65–71
de Melo J, Zibetti C, Clark BS, Hwang W, Miranda-Angulo AL, Qian J, Blackshaw S (2016) Lhx2
is an essential factor for retinal gliogenesis and Notch signaling. J Neurosci 36:2391–2405
Dittrich L, Morairty SR, Warrier DR, Kilduff TS (2015) Homeostatic sleep pressure is the primary
factor for activation of cortical nNOS/NK1 neurons. Neuropsychopharmacology 40:632–639
Dzierzewski JM, Dautovich N, Ravyts S (2018) Sleep and cognition in older adults. Sleep Med Clin
13:93–106
Eldred KC, Hadyniak SE, Hussey KA, Brenerman B, Zhang PW, Chamling X, Sluch VM, Welsbie
DS, Hattar S, Taylor J et al (2018) Thyroid hormone signaling specifies cone subtypes in human
retinal organoids. Science 362:eaau6348
Elmenhorst D, Meyer PT, Winz OH, Matusch A, Ermert J, Coenen HH, Basheer R, Haas HL,
Zilles K, Bauer A (2007) Sleep deprivation increases A1 adenosine receptor binding in the
human brain: a positron emission tomography study. J Neurosci 27:2410–2415
Elmenhorst D, Elmenhorst EM, Hennecke E, Kroll T, Matusch A, Aeschbach D, Bauer A (2017)
Recovery sleep after extended wakefulness restores elevated A1 adenosine receptor availability
in the human brain. Proc Natl Acad Sci USA 114:4243–4248
Ferreira JGP, Bittencourt JC, Adamantidis A (2017) Melanin-concentrating hormone and sleep.
Curr Opin Neurobiol 44:152–158
Fischer D, Lombardi DA, Marucci-Wellman H, Roenneberg T (2017) Chronotypes in the US—
influence of age and sex. PLoS One 12:e0178782
Foreman SW, Thomas KA, Blackburn ST (2008) Individual and gender differences matter in
preterm infant state development. J Obstet Gynecol Neonatal Nurs 37:657–665
Frank MG, Heller HC (1997) Development of diurnal organization of EEG slow-wave activity and
slow-wave sleep in the rat. Am J Phys 273:R472–R478
5 Winding the Clock: Development of Hypothalamic Structures Controlling. . . 125
Fu T, Towers M, Placzek MA (2017) Fgf10(+) progenitors give rise to the chick hypothalamus by
rostral and caudal growth and differentiation. Development 144:3278–3288
Gall AJ, Todd WD, Ray B, Coleman CM, Blumberg MS (2008) The development of day-night
differences in sleep and wakefulness in Norway rats and the effect of bilateral enucleation. J Biol
Rhythm 23:232–241
Galland BC, Taylor BJ, Elder DE, Herbison P (2012) Normal sleep patterns in infants and children:
a systematic review of observational studies. Sleep Med Rev 16:213–222
Hanne HM, Meadows JD, Trang C, Hereford B, Bharti K, Gorman MR, Mellon PL (2016) Deletion
of SIX3 or VAX1 in the SCN impairs circadian rhythms and fertility. In Endocrine Society’s
98th Annual Meeting and Expo, April 1–4, 2016, Boston
Hatori M, Gill S, Mure LS, Goulding M, O’Leary DD, Panda S (2014) Lhx1 maintains synchrony
among circadian oscillator neurons of the SCN. Elife 3:e03357
Hattar S, Liao HW, Takao M, Berson DM, Yau KW (2002) Melanopsin-containing retinal ganglion
cells: architecture, projections, and intrinsic photosensitivity. Science 295:1065–1070
Hrvatin S, Hochbaum DR, Nagy MA, Cicconet M, Robertson K, Cheadle L, Zilionis R, Ratner A,
Borges-Monroy R, Klein AM et al (2018) Single-cell analysis of experience-dependent
transcriptomic states in the mouse visual cortex. Nat Neurosci 21:120–129
Joseph D, Chong NW, Shanks ME, Rosato E, Taub NA, Petersen SA, Symonds ME, Whitehouse
WP, Wailoo M (2015) Getting rhythm: how do babies do it? Arch Dis Child Fetal Neonatal Ed
100:F50–F54
Jud C, Schmutz I, Hampp G, Oster H, Albrecht U (2005) A guideline for analyzing circadian wheel-
running behavior in rodents under different lighting conditions. Biol Proced Online 7:101–116
Lee JE, Wu SF, Goering LM, Dorsky RI (2006) Canonical Wnt signaling through Lef1 is required
for hypothalamic neurogenesis. Development 133:4451–4461
Liu J, Merkle FT, Gandhi AV, Gagnon JA, Woods IG, Chiu CN, Shimogori T, Schier AF, Prober
DA (2015) Evolutionarily conserved regulation of hypocretin neuron specification by Lhx9.
Development 142:1113–1124
Liu K, Kim J, Kim DW, Zhang YS, Bao H, Denaxa M, Lim SA, Kim E, Liu C, Wickersham IR et al
(2017) Lhx6-positive GABA-releasing neurons of the zona incerta promote sleep. Nature
548:582–587
Manning L, Ohyama K, Saeger B, Hatano O, Wilson SA, Logan M, Placzek M (2006) Regional
morphogenesis in the hypothalamus: a BMP-Tbx2 pathway coordinates fate and proliferation
through Shh downregulation. Dev Cell 11:873–885
Mariani J, Coppola G, Zhang P, Abyzov A, Provini L, Tomasini L, Amenduni M, Szekely A,
Palejev D, Wilson M et al (2015) FOXG1-dependent dysregulation of GABA/glutamate neuron
differentiation in autism spectrum disorders. Cell 162:375–390
Martinez-Jimenez CP, Eling N, Chen HC, Vallejos CA, Kolodziejczyk AA, Connor F, Stojic L,
Rayner TF, Stubbington MJT, Teichmann SA et al (2017) Aging increases cell-to-cell tran-
scriptional variability upon immune stimulation. Science 355:1433–1436
Maywood ES, Chesham JE, O’Brien JA, Hastings MH (2011) A diversity of paracrine signals
sustains molecular circadian cycling in suprachiasmatic nucleus circuits. Proc Natl Acad Sci
USA 108:14306–14311
Merkle FT, Maroof A, Wataya T, Sasai Y, Studer L, Eggan K, Schier AF (2015) Generation of
neuropeptidergic hypothalamic neurons from human pluripotent stem cells. Development
142:633–643
Miles LE, Dement WC (1980) Sleep and aging. Sleep 3:1–220
Mohawk JA, Takahashi JS (2011) Cell autonomy and synchrony of suprachiasmatic nucleus
circadian oscillators. Trends Neurosci 34(7):349–358
Morairty SR, Dittrich L, Pasumarthi RK, Valladao D, Heiss JE, Gerashchenko D, Kilduff TS (2013)
A role for cortical nNOS/NK1 neurons in coupling homeostatic sleep drive to EEG slow wave
activity. Proc Natl Acad Sci USA 110:20272–20277
Nelson AB, Faraguna U, Zoltan JT, Tononi G, Cirelli C (2013) Sleep patterns and homeostatic
mechanisms in adolescent mice. Brain Sci 3:318–343
126 D. W. T. Kim and S. Blackshaw
Newman EA, Kim DW, Wan J, Wang J, Qian J, Blackshaw S (2018a) Foxd1 is required for
terminal differentiation of anterior hypothalamic neuronal subtypes. Dev Biol 439:102–111
Newman EA, Wu D, Taketo MM, Zhang J, Blackshaw S (2018b) Canonical Wnt signaling
regulates patterning, differentiation and nucleogenesis in mouse hypothalamus and prethalamus.
Dev Biol 442:236–248
Ogawa K, Suga H, Ozone C, Sakakibara M, Yamada T, Kano M, Mitsumoto K, Kasai T, Kodani Y,
Nagasaki H et al (2018) Vasopressin-secreting neurons derived from human embryonic stem
cells through specific induction of dorsal hypothalamic progenitors. Sci Rep 8:3615
Ohyama K, Das R, Placzek M (2008) Temporal progression of hypothalamic patterning by a dual
action of BMP. Development 135:3325–3331
Parsons MJ, Brancaccio M, Sethi S, Maywood ES, Satija R, Edwards JK, Jagannath A, Couch Y,
Finelli MJ, Smyllie NJ et al (2015) The regulatory factor ZFHX3 modifies circadian function in
SCN via an AT motif-driven Axis. Cell 162:607–621
Partch CL, Green CB, Takahashi JS (2014) Molecular architecture of the mammalian circadian
clock. Trends Cell Biol 24:90–99
Pedersen NP, Ferrari L, Venner A, Wang JL, Abbott SBG, Vujovic N, Arrigoni E, Saper CB, Fuller
PM (2017) Supramammillary glutamate neurons are a key node of the arousal system. Nat
Commun 8:1405
Porter FD, Drago J, Xu Y, Cheema SS, Wassif C, Huang SP, Lee E, Grinberg A, Massalas JS,
Bodine D et al (1997) Lhx2, a LIM homeobox gene, is required for eye, forebrain, and definitive
erythrocyte development. Development 124:2935–2944
Qian X, Nguyen HN, Song MM, Hadiono C, Ogden SC, Hammack C, Yao B, Hamersky GR,
Jacob F, Zhong C et al (2016) Brain-region-specific organoids using mini-bioreactors for
modeling ZIKV exposure. Cell 165:1238–1254
Qian X, Jacob F, Song MM, Nguyen HN, Song H, Ming GL (2018) Generation of human brain
region-specific organoids using a miniaturized spinning bioreactor. Nat Protoc 13:565–580
Reddy AB, O’Neill JS (2009) Healthy clocks, healthy body, healthy mind. Trends Cell Biol
20:36–44
Roenneberg T, Wirz-Justice A, Merrow M (2003) Life between clocks: daily temporal patterns of
human chronotypes. J Biol Rhythm 18:80–90
Romanov RA, Zeisel A, Bakker J, Girach F, Hellysaz A, Tomer R, Alpar A, Mulder J, Clotman F,
Keimpema E et al (2017) Molecular interrogation of hypothalamic organization reveals distinct
dopamine neuronal subtypes. Nat Neurosci 20:176–188
Roy A, de Melo J, Chaturvedi D, Thein T, Cabrera-Socorro A, Houart C, Meyer G, Blackshaw S,
Tole S (2013) LHX2 is necessary for the maintenance of optic identity and for the progression of
optic morphogenesis. J Neurosci 33:6877–6884
Sack RL, Auckley D, Auger RR, Carskadon MA, Wright KP Jr, Vitiello MV, Zhdanova IV,
American Academy of Sleep, M (2007) Circadian rhythm sleep disorders: part I, basic
principles, shift work and jet lag disorders. An American Academy of Sleep Medicine review.
Sleep 30:1460–1483
Saito YC, Tsujino N, Hasegawa E, Akashi K, Abe M, Mieda M, Sakimura K, Sakurai T (2013)
GABAergic neurons in the preoptic area send direct inhibitory projections to orexin neurons.
Front Neural Circuits 7:192
Salvatierra J, Lee DA, Zibetti C, Duran-Moreno M, Yoo S, Newman EA, Wang H, Bedont JL, de
Melo J, Miranda-Angulo AL et al (2014) The LIM homeodomain factor Lhx2 is required for
hypothalamic Tanycyte specification and differentiation. J Neurosci 34:16809–16820
Saper CB, Fuller PM, Pedersen NP, Lu J, Scammell TE (2010) Sleep state switching. Neuron
68:1023–1042
Scammell TE, Arrigoni E, Lipton JO (2017) Neural circuitry of wakefulness and sleep. Neuron
93:747–765
Shimogori T, Lee DA, Miranda-Angulo A, Yang Y, Wang H, Jiang L, Yoshida AC, Kataoka A,
Mashiko H, Avetisyan M et al (2010) A genomic atlas of mouse hypothalamic development.
Nat Neurosci 13:767–775
5 Winding the Clock: Development of Hypothalamic Structures Controlling. . . 127
VanDunk C, Hunter LA, Gray PA (2011) Development, maturation, and necessity of transcription
factors in the mouse suprachiasmatic nucleus. J Neurosci 31:6457–6467
Venner A, Anaclet C, Broadhurst RY, Saper CB, Fuller PM (2016) A novel population of wake-
promoting GABAergic neurons in the ventral lateral hypothalamus. Curr Biol 26:2137–2143
Wang L, Meece K, Williams DJ, Lo KA, Zimmer M, Heinrich G, Martin Carli J, Leduc CA, Sun L,
Zeltser LM et al (2015) Differentiation of hypothalamic-like neurons from human pluripotent
stem cells. J Clin Invest 125:796–808
Wataya T, Ando S, Muguruma K, Ikeda H, Watanabe K, Eiraku M, Kawada M, Takahashi J,
Hashimoto N, Sasai Y (2008) Minimization of exogenous signals in ES cell culture induces
rostral hypothalamic differentiation. Proc Natl Acad Sci USA 105:11796–11801
Weber F, Dan Y (2016) Circuit-based interrogation of sleep control. Nature 538:51–59
Wilcox AG, Vizor L, Parsons MJ, Banks G, Nolan PM (2017) Inducible knockout of mouse Zfhx3
emphasizes its key role in setting the pace and amplitude of the adult circadian clock. J Biol
Rhythm 32:433–443
Wolf A, Ryu S (2013) Specification of posterior hypothalamic neurons requires coordinated
activities of Fezf2, Otp, Sim1a and Foxb1.2. Development 140:1762–1773
Xie Y, Kaufmann D, Moulton MJ, Panahi S, Gaynes JA, Watters HN, Zhou D, Xue HH, Fung CM,
Levine EM et al (2017) Lef1-dependent hypothalamic neurogenesis inhibits anxiety. PLoS Biol
15:e2002257
Yamaguchi Y, Suzuki T, Mizoro Y, Kori H, Okada K, Chen Y, Fustin JM, Yamazaki F,
Mizuguchi N, Zhang J et al (2013) Mice genetically deficient in vasopressin V1a and V1b
receptors are resistant to jet lag. Science 342:85–90
Yao L, Liu Y, Qiu Z, Kumar S, Curran JE, Blangero J, Chen Y, Lehman DM (2017) Molecular
profiling of human induced pluripotent stem cell-derived hypothalamic neurones provides
developmental insights into genetic loci for body weight regulation. J Neuroendocrinol 29(2)
Pituitary Development and Organogenesis:
Transcription Factors in Development 6
and Disease
Abstract
Many transcription factors that regulate pituitary gland development and specifi-
cation of hormone-producing cell types have been discovered. Here we focus on
13 transcription factors that play key roles in these processes, many of which are
mutated in patients with congenital pituitary insufficiency disorders. We present
the methods by which each of these transcription factors were identified. Each of
the genes we discuss has been well studied in genetically engineered mice,
facilitating an understanding of the impact of each single gene defect on the
expression of downstream targets. Finally, we highlight technical advances that
promise to yield future insights into epigenetic regulation of cell fate.
Keywords
Cell specification · Stem cell · Hypopituitarism
6.1 Introduction
The pituitary gland is located at the base of the brain, and it is a small organ,
approximately the size of a pea in humans. It is often referred to as the master
gland because it is a key regulator of multiple organ systems, controlling growth,
fertility, stress response, and homeostasis. The pituitary gland develops between
5 and 9 weeks of gestation in humans and embryonic day 10–19 in mice (Xue et al.
2013). An invagination of oral ectoderm, called Rathke’s pouch, forms the anterior
and intermediate lobes of the pituitary gland, and evagination of the neural ectoderm
forms the posterior lobe of the pituitary gland and pituitary stalk. The posterior lobe
contains the axon terminals for oxytocin and vasopressin and pituicytes. The anterior
pineal
gland
cerebellum
cortex
midbrain
thalamus
olfactory
blub hypothalamus medulla
pons
pituitary
gland
supraoptic
nucleus
hypothalamus
paraventricular
nucleus
optic
chiasm
posterior lobe
oxytocin,
pituitary vasopressin
stalk
Fig. 6.1 Anatomy of the pituitary gland and base of the hypothalamus. The pituitary gland
contains anterior (ant) and intermediate (int) lobes derived from the oral ectoderm that forms
Rathke’s pouch. The posterior lobe (post) and pituitary stalk are derived from neural ectoderm of
the ventral diencephalon, and they contain the axon terminals of vasopressin and oxytocin neurons
that project from both the supraoptic nucleus and the paraventricular nucleus of the hypothalamus.
Hypothalamic peptides regulate the release of anterior pituitary hormones into the hypophyseal
portal system [adapted from Rappaport and Amselem (2001)]
The hypophyseal portal system refers to the blood vessels that provide connections
between the hypothalamus and the anterior pituitary gland. This vasculature
develops e14.5–e18.5 in mice and appears complete by 12 weeks of gestation in
humans (Xue et al. 2013). The capillaries are highly fenestrated, facilitating rapid
molecular exchange between the hypothalamus and pituitary lobes. Oxygen levels
and blood flow are regulated and have an effect on pituitary hormone production and
secretion (Lafont et al. 2010; Schaeffer et al. 2010).
The mature pituitary gland has six distinct hormone-producing cell types, as well
as folliculo-stellate cells and pituicytes (Fig. 6.2). The hormone-producing cell types
in the anterior lobe include corticotropes that produce adrenocorticotropin (ACTH)
from the pro-hormone pro-opiomelanocortin (POMC), somatotropes that produce
GH, lactotropes that produce prolactin, gonadotropes that produce the heterodimeric
glycoprotein hormones luteinizing hormone (LH) and follicle-stimulating hormone
(FSH) and thyrotropes that produce the heterodimeric glycoprotein hormone thy-
roid-stimulating hormone (TSH). The intermediate lobe, which remains distinct in
rodents, contains melanotropes that produce melanocyte-stimulating hormone from
POMC. Folliculo-stellate (FS) cells are non-endocrine cells that were named because
of the star-like morphology of cytoplasmic processes (Rinehart and Farquhar 1953).
FS cells appear between postnatal days 10 and 20 in the rat, throughout the anterior
lobe parenchyma, directly adjacent to the hormone-producing cells (Soji et al. 1997).
FS cells serve a support function for hormone-producing cells through the release of
cytokines and growth factors. FS cells are excitatory and are involved in the
coordinated, pulsatile release of hormones from the hormone-producing cells,
which themselves form homotypic networks (Fauquier et al. 2001; Mollard et al.
2012). Pituicytes are glial-like cells located within the posterior lobe. The larger
parenchymatous pituicytes regulate hormone output by completely enveloping the
neurosecretory processes of the oxytocin and vasopressin axons when hormone
secretion is low, and they recede as hormone secretion increases. They also regulate
hormone output by secretion of taurine (Rosso and Mienville 2009). The smaller
fibrous pituicytes do not appear to be involved in this regulatory function and are less
well understood (Pow et al. 1989).
The vasculature is important for pituitary gland function because it transports
molecules from pituitary target organs that feedback at the level of the pituitary gland
132 A. Z. Daly and S. A. Camper
Corticotrope
Melanotrope
TPit Pax7
Somatotrope
Prop1 Pou1f1
Lactotrope
SOX2 Thyrotrope
stem cell
SF1
Gonadotrope
pituitary population
Fraction of total
Proliferating
Precursors
Quiescent
Precursors
Differentiated
Cells
E9.5 E11.5 E13.5 E17.5
Fig. 6.2 Critical transcription factors control the development of specialized hormone-producing
cells. The major hormone-producing cell types of the anterior and intermediate lobes of the pituitary
gland are derived from SOX2-expressing progenitor cells. Combinations of transcription factors
drive specific cell fates and antagonize differentiation into alternate fates. Each cell type has
distinctive secretory granules and shape (Matsumoto and Ishii 1992). Proliferating precursors
leave the cell cycle and differentiate into hormone-producing cells during gestation
Some of the major signaling pathways that have been implicated in pituitary devel-
opment are SHH, FGF, BMP, Wnt, Notch, Hippo, EGF, and retinoic acid (Cohen
et al. 1999; Djakoure et al. 1996; Lodge et al. 2019; Zhu et al. 2007). Hypothalamic
SHH is necessary for inducing expression of the pituitary transcription factors LHX3
and LHX4, which drive progenitor proliferation and prevent cell death (Carreno et al.
6 Pituitary Development and Organogenesis: Transcription Factors in. . . 133
2017). FGF8 and FGFR1 mutations cause pituitary hormone deficiency in humans
(Fang et al. 2016), and mouse studies show reduced cell proliferation and enhanced
cell death in the pituitary primordia of Fgf8, Fgf10, and Fgfr2 mutants (Ohuchi et al.
2000). BMP and the antagonist noggin affect pituitary growth and shape during
development, and altered BMP signaling is associated with pituitary adenomas
(Davis and Camper 2007; Ericson et al. 1998; Giacomini et al. 2006; Treier et al.
2001). Similarly, WNT signaling regulates pituitary organogenesis and acts in a
paracrine manner to stimulate excess cell proliferation in adenomas (Chambers
et al. 2013). The ligand, beta-catenin, is a cofactor for several key pituitary transcrip-
tion factors including PITX2, TCF7L2 (TCF4), LEF1, NR5A1, and PROP1. Notch
signaling regulates the timing of pituitary progenitor cell cycle exit and differentiation
(Cheung et al. 2018a; Raetzman et al. 2004; Zhu et al. 2006). YAP and TAZ are
transcriptional regulators downstream of the Hippo signaling pathway, which
suppresses their function by phosphorylation. This pathway plays an essential role
in the regulation of pituitary progenitor expansion and normal organ size (Lodge et al.
2019). There is a great deal of cross talk between these pathways, in that perturbation
of one signaling pathway typically affects the other signaling pathways. EGF and
retinoic acid signaling are recognized as important signaling pathways but their
precise roles during organogenesis have not been established (Rhodes et al. 1993;
Roh et al. 2001).
6.1.3 Hypopituitarism
The major pathologies associated with the pituitary are adenomas, which can be
associated with either over- or underproduction of hormones, and congenital or
acquired hypopituitarism (Higham et al. 2016; Melmed 2011). The main emphasis
of this review will be on congenital hypopituitarism, which had an estimated
prevalence of 45.5 per 100,000 people (Regal et al. 2001). This is a genetically
heterogeneous condition that can present with a single pituitary hormone deficiency
or with multiple deficiencies. About 50% of cases that originally present with
isolated growth hormone deficiency (IGHD) progress to combined pituitary hor-
mone deficiency (CPHD). CPHD is defined as a reduction of at least two pituitary
hormones, and usually, GH is one of them. It has a prevalence of 1 in 8000
individuals, and more than 30 genes have been implicated as causal factors for
CPHD (Fang et al. 2016).
6.1.4 Overview
Here we review the transcription factors that are critical for developing the specialized
cells of the anterior and intermediate lobes of the pituitary gland, emphasizing the
history of scientific discovery surrounding each factor, from its discovery through to
its most recent characterization. This reveals the continuing evolution and diversity of
effective technical approaches. These include identification of cis-acting sequences
134 A. Z. Daly and S. A. Camper
important for regulation of hormone gene expression and identification of the trans-
acting factors that bind those sites, positional cloning of mutations in mice and human
patients, and exploration of epigenetic regulation of chromatin accessibility. We also
highlight areas for future discovery that may take advantage of high-throughput
sequencing and high-efficiency, targeted germline disruption tools.
Sox2 was first discovered in a whole embryo cDNA screen in search of genes similar
to the male, sex-determining gene Sry, which binds DNA through a motif or box
characteristic of high mobility group proteins (Gubbay et al. 1990). Sox2 is
expressed in progenitor cell populations in many tissues, and its expression is
negatively correlated with differentiation (Avilion et al. 2003; Li et al. 1998;
Zappone et al. 2000). SOX2 is one of the factors that, along with OCT3/4,
c-MYC, and KLF4, were demonstrated by Yamanaka and colleagues to be sufficient
to induce pluripotency in differentiated cells from mice and humans (Takahashi et al.
2007; Takahashi and Yamanaka 2006).
Sox2 is expressed in the developing pituitary and hypothalamus, and it has a role in
pituitary development (Takahashi et al. 2007; Takahashi and Yamanaka 2006).
Embryos homozygous for Sox2 loss of function alleles die around implantation,
and heterozygous mice have genetic background-dependent reduction in body size
and male infertility (Avilion et al. 2003). The developing pituitaries of heterozygous
mutants are bifurcated, similar to the dysmorphology observed in Wnt5a mutants,
and both males and females exhibit reduced differentiation into somatotropes and
gonadotropes (Fig. 6.3) (Cha et al. 2004; Rizzoti and Lovell-Badge 2005). About
two-thirds of the heterozygotes survive to adulthood, and in these survivors, the
reduction in pituitary size is modest, and GH and LH content is only significantly
reduced in males.
Martinez-Barbera and colleagues generated a conditional deletion of Sox2 using
the Hesx1-cre strain, deleting Sox2 in the pituitary gland and areas of the brain
between e10.5 and e12.5 (Rizzoti and Lovell-Badge 2005). This resulted in severe
pituitary hypoplasia and greatly reduced the expressions of Pou1f1, Gh, and Tshb.
Gonadotropin deficiency appeared to be secondary to the reduction in GnRH
neurons in these mice. Cell proliferation is significantly decreased in the condition-
ally deleted Sox2 pituitary glands, and it appears that the progenitors have not
migrated from the progenitor zone of Rathke’s pouch to colonize the anterior lobe.
This suggests that Sox2 has an important role in pituitary progenitor proliferation and
transition to differentiation.
6 Pituitary Development and Organogenesis: Transcription Factors in. . . 135
Fig. 6.3 The effect of transcription factor mutations on pituitary development has been revealed
using genetically engineered and spontaneous mutant mice. Deletion of Pitx1 has modest effects
including slight reduction of TSH and LH, slight increase in POMC and no change in GH and
αGSU, as observed by in situ hybridization in P0 animals (Szeto et al. 1999). In situ hybridization in
136 A. Z. Daly and S. A. Camper
Robinson and colleagues discovered that Sox2-expressing cells in the adult mouse
pituitary were still mitotically active and that, given appropriate culture conditions,
these SOX2-positive cells can self-renew and differentiate in vitro into all of the
hormone-producing cells of the anterior lobe of the pituitary gland (Fauquier et al.
2008). Both Lovell-Badge and Martinez-Barbera and colleagues used lineage tracing
to demonstrate that SOX2-positive cells give rise to all of the different hormone-
producing cells in the pituitary gland in vivo (Andoniadou et al. 2013; Rizzoti et al.
2013). SOX2-positive cells contribute to tissue homeostasis, and they are involved in
responding to end-organ ablation, such as adrenalectomy and gonadectomy. Mice
with a constitutively active allele of β-catenin exhibit clusters of pituitary cells that
overexpress β-catenin, express Sox2, and induce transformation of neighboring cells
that mimic craniopharyngioma, a childhood tumor (Gaston-Massuet et al. 2011).
Taken together these studies established that SOX2-expressing cells represent
progenitors of the hormone-producing cells in the developing pituitary.
Most patients with SOX2 mutations have serious eye abnormalities that can include
anophthalmia, microphthalmia, and/or optic nerve hypoplasia, and hypopituitarism
characterized by pituitary gland hypoplasia and hypogonadotropic hypogonadism
Fig. 6.3 (continued) newborn mice homozygous for a hypomorphic allele of Pitx2, Pitx2neo/neo,
revealed normal Pomc expression, reduction in Gh and Tshb expression, and no detectable Lhb or
Fshb transcripts (panels bordered in purple) (Suh et al. 2002). Pitx2/ embryos had arrested the
development of Rathke’s pouch and reduced expression of αGSU at e13.5 mice (Lin et al. 1999).
Trace amounts of Pomc transcripts were present, but all other hormones were not detectable (not
shown). In situ hybridization in the Lhx3 knockout mice revealed a near-total loss of Pomc, Tshb,
Gh, and Cga (αGSU) transcripts, and immunohistochemistry failed to detect significant amounts of
LH. In situ hybridization of Lhx4 knockout pituitaries revealed a significant reduction in Tshb, Lhb,
Gh, and Cga transcripts at e18.5 (Sheng et al. 1997), and immunohistochemistry showed largely
similar POMC expression in e14.5 mice (Sheng et al. 1997). Immunohistochemistry of Sox2
heterozygous null mice revealed reduced staining for GH and LH at e18.5 (Kelberman et al.
2006). There is no detectable Pou1f1 staining in Hesx cre/+; Sox2flox/flox mice at e14.5 or later in
gestation (not shown) (Jayakody et al. 2012). The staining for αGSU and POMC was somewhat
normal at e15.5, considering the hypoplasia of the organ. Only Pou1f1 independent
TSH-expressing cells were detected. Immunohistochemistry of Prop1/ pituitaries revealed
POMC expression but little or no detectable caudomedial staining for TSH, LH, or GH at e18.5
(Nasonkin et al. 2004). αGSU staining was modestly reduced at e14.5 in Prop1df/df mice, but
NR5A1 cells were normal at birth (Raetzman et al. 2002). In situ hybridization analysis of e17.5
Hesx1R160C/R160C mutants revealed approximately normal levels of Pomc, Tshb, Lhb, Gh, and Cga
expression, although the pituitaries were dysmorphic (Sajedi et al. 2008). These missense mutant
mice have similar pituitary features to mice homozygous for a null allele, but the forebrain is less
affected. Analysis of newborn SF1-mutant mice revealed undetectable LH, diminished Cga
transcripts and normal immunostaining for ACTH, TSH, and GH (Ingraham et al. 1994).
Pou1f1dw/dw mutant newborns have normal POMC immunostaining, undetectable TSH and GH,
and increased LH (Mortensen et al. 2011). Cga appeared reduced. Immunohistochemical analysis
of Tpit knockouts revealed undetectable POMC, no effect on caudomedial staining of GH, TSH,
LH, or αGSU in the anterior lobe, but ectopic expression of TSH, LH, and αGSU in the intermediate
lobe (Pulichino et al. 2003b).
6 Pituitary Development and Organogenesis: Transcription Factors in. . . 137
(Gaston-Massuet et al. 2011; Gubbay et al. 1990; Kelberman et al. 2006; Ragge et al.
2005). While all of the patients with SOX2 mutations exhibited reduced gonadotrope
function, the degree of GH deficiency varied. The response of patients to GnRH
stimulation was also variable.
6.3 PITX Factors (PITX1: Ptx1, Bft, P-OTX; PITX2: Brx, Munc30,
Otlx2, Rieg)
The PITX transcription factors, PITX1 and PITX2, are expressed early in pituitary
gland development and persist through adulthood. Their names derive from their
pituitary expression and paired-like homeobox DNA-binding domains, and they
have important functions in the hind limb, heart, eye, and other tissues (Gage et al.
1999; Kitamura et al. 1999; Lin et al. 1999; Lu et al. 1999; Szeto et al. 1999). These
factors were found in very different ways, and they have overlapping functions
within the pituitary gland (Charles et al. 2005).
Pitx1 was discovered by searching for factors that bind the regulatory elements of
the pro-opiomelanocortin gene (Pomc). POMC is a pro-hormone that is expressed in
both corticotropes of the anterior lobe and melanotropes or the intermediate lobe.
POMC undergoes differential cleavage to produce adrenocorticotropin (ACTH) in
corticotropes and melanocyte-stimulating hormone (MSH) and β-endorphin in
melanotropes. In vitro analysis revealed that the major regulatory region of the Pomc
promoter-proximal region was between 166 and 480 bp upstream of the transcription
start site (Therrien and Drouin 1991). Drouin and colleagues found that deleting a small
section from the middle of this element nearly ablated Pomc expression (Lamonerie
et al. 1996). They screened a cDNA expression library from AtT-20 cells, an
immortalized corticotrope-like cell, for factors that bind this element, and identified
Pitx1, which has high homology with the mouse Otx genes. Pitx1 was identified
independently by screening the same cDNA expression library for factors that interact
with POU1F1 (Szeto et al. 1996), a pituitary transcription factor described below.
PITX2 was discovered as one cause of Rieger syndrome, which was first
described in the 1800s as a genetic disorder characterized by eye malformations,
dental hypoplasia, craniofacial dysmorphism, and umbilical stump abnormalities.
The causal gene was mapped to a 50-kb region on chromosome 4 (Datson et al.
1996). Genomic DNA from CpG islands in this region was used to screen human
craniofacial and fetal brain cDNA libraries, and a contig of RIEG cDNAs was
obtained (Semina et al. 1996). RIEG was most closely related to Pitx1, and it was
later renamed PITX2. Murray and colleagues reported mutations in six families with
Rieger syndrome. Pitx2 was discovered independently by screening an adult mouse
pituitary cDNA library for homeobox genes (Gage and Camper 1997). It was
predicted to be a regulator of anterior structure formation in mice based on its
expression pattern and its location on mouse chromosome 3 within a region of
synteny homology to the region of human 4q25 where Rieger syndrome had been
mapped.
138 A. Z. Daly and S. A. Camper
The expression patterns of Pitx1 and Pitx2 differ, but there are some regions of
striking overlap. Pitx1 is expressed in the pituitary, intestine, tongue, oral epithelium,
the mandible, salivary glands, duodenum, nasal epithelium, and the hind limb
(Lamonerie et al. 1996; Lanctot et al. 1999; Szeto et al. 1996). Pitx1 is expressed
in the pituitary primordia by embryonic day 9.5 (e9.5) (Treier et al. 1998). Pitx2 is
expressed in the eye, the dental lamina, limb mesenchyme, umbilical vessels, and in
Rathke’s pouch as early as e8.5 (Semina et al. 1996). The early expression of both
PITX factors suggested their importance for pituitary development.
PITX1 binds the promoters of several genes that encode pituitary hormones, includ-
ing Cga (chorionic gonadotropin alpha, the common subunit of the heterodimeric
hormones LH, FSH, and TSH), Gh, and Prl, and induces their transcription (Szeto
et al. 1996). PITX1 was also implicated in the transcription of Gnrhr, Pou1f1, Tshb,
Nr5a1, and Lhx3 using cell transfection assays (Tremblay et al. 1998). However,
immunostaining suggests that gonadotropes and thyrotropes have the most abundant
expression of PITX1 and PITX2 in the adult pituitary gland (Charles et al. 2006).
PITX2 activity is enhanced by β-catenin, and the complex regulates Cyclin D2
expression (Kioussi et al. 2002).
Mice homozygous for Pitx1 deletion had disrupted hind limb formation and a very
mild pituitary phenotype with a slight reduction of gonadotropes and thyrotropes
(Szeto et al. 1999). All of the major transcription factors were largely unaffected, and
Pomc expression was normal.
By comparison, Pitx2 had a much stronger effect on the developing pituitary.
Four independent Pitx2 knockout mouse lines were published in 1999 (Gage et al.
1999; Kitamura et al. 1999; Lin et al. 1999; Lu et al. 1999). Camper and colleagues
developed several alleles, which included a hypomorphic or reduced function allele,
Pitx2neo, a null allele, Pitx2, and an inducible null allele (Suh et al. 2002). Various
combinations of these alleles were used to assess dosage-sensitive effects of Pitx2
loss of function, i.e., an allelic series. Pitx2/ and Pitx2neo/neo embryos die in utero,
at e13.5 and e18.5, respectively. Both exhibit right lung isomerism, and dosage-
sensitive effects on eye, pituitary gland, and heart development. The ventral body
wall of Pitx2/ embryos fails to close, resulting in the externalization of the heart,
liver, and abdominal organs. Pitx2 heterozygotes accurately model the ocular
features of Rieger syndrome (Chen and Gage 2016).
The initial induction of Rathke’s pouch occurs at e10.5 in Pitx2/ embryos, but
there are extensive cell death and reduced cell proliferation, resulting in a very small
pituitary primordium at e12.5 (Charles et al. 2005; Gage et al. 1999). Hesx1, Prop1,
6 Pituitary Development and Organogenesis: Transcription Factors in. . . 139
and Lhx4 are not expressed in these mutants, but Lhx3 is activated. Little or no
hormone cell specification is detected (Lin et al. 1999). In contrast to the small
pituitaries in Pitx2/ and Pitx2neo/ embryos, there was no obvious reduction in the
size of Pitx2neo/neo pituitaries (Suh et al. 2002). Somatotropes and thyrotropes were
reduced, and gonadotropes were ablated. Pitx2neo/neo pituitaries had reduced expres-
sion of the transcription factors Gata2, Egr1, Nr5a1, and Pou1f1.
The overlapping expression patterns of Pitx1 and Pitx2 suggested the possibility
of compensatory function. Indeed, double heterozygotes (Pitx1+/, Pitx2+/ and
Pitx1+/, Pitx2+/neo) have very poor viability, and hypoplastic pituitaries (Charles
et al. 2005). Lhx3 expression was not detectable in rare double mutants, suggesting
that Pitx genes are required for Lhx3 expression.
Heterozygous PITX1 loss of function mutations have been reported in two families
with lower extremity abnormalities (Alvarado et al. 2011; Gurnett et al. 2008;
Klopocki et al. 2012). Genomic rearrangements that result in ectopic expression of
PITX1 in the developing forelimbs cause Liebenberg syndrome, a homeotic trans-
formation of the arms to a leg morphology (Kragesteen et al. 2019; Spielmann et al.
2012).
Rieger syndrome is a rare, autosomal dominant disorder affecting the develop-
ment of the eyes, teeth, and abdominal wall. It can be caused by mutations in PITX2
or FOXC1. The majority of patients with Rieger syndrome have normal height,
although there are some anecdotal reports of Rieger syndrome patients with growth
insufficiency. For example, a three-generation family with Rieger syndrome and
either short stature or growth hormone deficiency was reported (Sadeghi-Nejad and
Senior 1974). The proband responded to GH replacement therapy. The nature of the
mutation in this family and other historical cases is unknown. Many Rieger syn-
drome patients have been screened for PITX2 mutations, but no systematic screening
of CPHD patients without eye defects has been reported.
LHX proteins contain both LIM-domains, composed of two contiguous zinc fingers
and a Homeobox-domain. Three LHX factors involved in pituitary development,
Lhx2, Lhx3, and Lhx4, and they were discovered independently. Lhx2 was discov-
ered in a screen of cell lines for VDJ recombinase activity (Xu et al. 1993). Lhx3 was
discovered in a screen for Lhx factors in a cDNA library generated from Xenopus
gastrula (called Xlim3 in Xenopus) (Taira et al. 1992). Lhx4 (previously known as
Gsh-4) was discovered in a screen for homeobox genes in mouse genomic DNA
(Singh et al. 1991). Lhx2 is expressed in developing B cells, telencephalon, retina,
140 A. Z. Daly and S. A. Camper
Mouse knockouts of Lhx2, Lhx3, and Lhx4 revealed the importance of each gene for
pituitary development (Ellsworth et al. 2008; Li et al. 1994; Raetzman et al. 2002;
Sheng et al. 1996, 1997; Zhao et al. 2010).
Lhx2 is expressed in the developing ventral diencephalon from e9.5 to e12.5 in
regions that become the pituitary infundibulum and posterior lobe (Zhao et al. 2010).
Embryos homozygous for Lhx2 null alleles have pituitary dysmorphology including
bifurcation of Rathke’s pouch, excess proliferation in the infundibular area, and
failure of the neural ectoderm to evaginate. Despite this, all hormone-producing cell
types of the anterior lobe are detectable, consistent with the expression of critical
signaling molecules such as BMP and FGFs by the abnormal infundibular structure.
The mice are not viable and also have anophthalmia, anemia, and malformations of
the cortex (Porter et al. 1997). A cohort of 59 patients with pituitary deficiency and
eye abnormalities were negative for LHX2 mutations (Perez et al. 2012).
Lhx3 is expressed in the pituitary placode at e9.5 and persists through adulthood
(Sheng et al. 1996). Lhx3+/ mice are phenotypically normal and fertile, but mice
homozygous for the null allele were either stillborn or died within 24 h after birth.
Rathke’s pouch formed in Lhx3/ embryos but failed to expand. The thin, under-
developed pouch fails to separate from the oral ectoderm. These mutant pouches
exhibited reduced cell proliferation and enhanced cell death, and there is evidence of
abnormal dorsal–ventral polarity (Ellsworth et al. 2008). The cyclin-dependent
kinase inhibitor Cdkn1a (p21) is expressed more dorsally than normal, and
Cdkn1c (p57) is absent in the caudal region of the pouch (Gergics et al. 2015).
Hesx1 and Isl1 expression were initiated at e10.5 but failed to be sustained at e12.5.
Induction of lineage-specific transcription factor gene expression—Tbx19, Neurod1,
Nr5a1, and Pou1f1—were very poor. Lhx3/ embryos contained a few
differentiated corticotropes but no other differentiated hormone-producing cells
were detected.
Lhx4 is expressed in the pituitary primordium between e9.5 and e12.5, and it is
restricted to the prospective anterior lobe at e12.5 (Sheng et al. 1997). Expression
diminishes at e15.5, but it is detectable in adult pituitary anterior and intermediate
lobes. Lhx4/ newborns appear normal but die shortly after birth due to failure of
the lungs to inflate (Li et al. 1994). Dexamethasone treatment, used to accelerate
lung development, improved but did not completely rescue survival. Pituitary
insufficiency can result in poor lung development in certain genetic backgrounds
(Nasonkin et al. 2004). Lhx4/ embryos have anterior pituitary lobe hypoplasia
(Sheng et al. 1997). Enhanced cell death is evident at e12.5, and Lhx3 expression is
significantly delayed (Gergics et al. 2015; Raetzman et al. 2002). Isl1 expression
6 Pituitary Development and Organogenesis: Transcription Factors in. . . 141
does not become ventralized at e11.5 and it is not detected at e12.5. Pitx2 expression
is also not maintained. Expression of pituitary hormone genes is reduced but
detectable.
The functions of Lhx3 and Lhx4 overlap (Sheng et al. 1997). Both Lhx3+/,
Lhx4/ and Lhx3/, Lhx4+/ embryos can form Rathke’s pouch, but it fails to
expand and is more severely affected than in either single mutant. Lhx3/, Lhx4/
double mutants have a very severe pituitary phenotype (Sheng et al. 1997). The
double mutants form a pouch rudiment, and invagination of the oral ectoderm
occurs, but the rudiment fails to separate from the oral ectoderm and remains at
the pharyngeal cavity beneath the palate.
Hesx1 was discovered by a screen for Homeobox factors in an Embryonic Stem cell
cDNA library, and independently, in a screen for homeobox factors expressed
during gastrulation, at e7.5 (Hermesz et al. 1996; Thomas et al. 1995; Thomas and
Rathjen 1992). Hesx1 is expressed sequentially in the anterior visceral endoderm,
anterior definitive endoderm, and the cephalic neural plate. While officially known
142 A. Z. Daly and S. A. Camper
as Hesx1, it was also named Rathke’s Pouch homeobox (Rpx1) because of its robust
expression in the developing pituitary primoridia at e9.5. Its expression in the
developing pituitary gland is extinguished by PROP1 at e14.5 (Dasen et al. 2001;
Gage et al. 1996).
Robinson and colleagues generated a Hesx1 deletion in mice and demonstrated that
93% of homozygous mutants died before weaning (Dattani et al. 1998). The
phenotype was variable and included developmental abnormalities of the forebrain,
eye, and pituitary gland. The pituitary glands were missing (5%), misplaced, dys-
morphic, and/or small (Dasen et al. 2001). Hesx1 deficiency permits excessive FGF
signaling, resulting in multiple invaginations of oral ectoderm instead of a single
Rathke’s pouch (Dasen et al. 2001). Most mice heterozygous for the deletion were
normal, but 1% displayed mild abnormalities of the same organs affected in
homozygous mice.
Rosenfeld and colleagues demonstrated that HESX1 acts as a repressor of Pou1f1
expression by binding the Pou1f1 early enhancer (Dasen et al. 2001; Olson et al.
2006). HESX1 binds the co-repressor transducin-like enhancer of split 1 (TLE1)
through its engrailed homology domain and also interacts with a corepressor, histone
deacetylase complex containing Nuclear Receptor Co-Repressor (NCoR). Forced
expression of HESX1 and TLE1 during pituitary development is sufficient to
suppress POU1F1-dependent cell lineages. Both HESX1 and PROP1 are paired-
type homeodomain transcription factors that can compete for the same DNA binding
sites. In the presence of beta-catenin, PROP1 binds the early enhancer of Pou1f1 and
initiates Pou1f1 gene expression.
Both Lhx3 and Pitx2 are necessary upstream regulators of Hesx1 in the develop-
ing pituitary gland (Gage et al. 1999; Sheng et al. 1996). A 100-bp element that
directs spatial expression of Hesx1 in Rathke’s pouch lies ~3.3 kb 30 of the gene and
is bound by PITX2 and GATA2 (Chou et al. 2006). The promoter proximal
570 bp, exons 1 and 2 and intron 1 of Hesx1 are sufficient for transgene reporter
expression in the anterior visceral endoderm, anterior neural ectoderm, and anterior
neural plate. A negative element that suppresses Hesx1 expression in the hypothala-
mus, and responds to inductive signals from Rathke’s pouch, has been mapped 568
to 532 bp upstream of the gene.
The phenotypes of Hesx1 knockout mice suggested that HESX1 mutations might
cause pituitary and anterior structure malformations in humans. A screen of
61 patients with holoprosencephaly, Septo-Optic Dysplasia (SOD), or pituitary
insufficiency uncovered two siblings with agenesis of the corpus callosum and
6 Pituitary Development and Organogenesis: Transcription Factors in. . . 143
6.5.4 Summary
Hesx1 was discovered in two independent cDNA screens for homeobox transcription
factors. LHX1, LHX3, and LMX1B activate its expression in early head development,
and PITX2 and GATA2 activate its expression in Rathke’s pouch. Pituitary expres-
sion is extinguished by PROP1, which also releases HESX1-mediated repression of
Pou1f1, activating the development of somatotropes, lactotropes, and thyrotropes.
Homozygous mutations in mice and heterozygous mutations in humans cause variable
phenotypes that include reduced forebrain, anophthalmia, and pituitary dysfunction.
The cause of variable presentation is not known.
6.6 PROP1
narrowed to a 400–600 kb window (Sornson et al. 1996). Genomic clones from this
region were probed with cDNA from the developing pituitary, resulting in the
identification of Prop1, which encodes a 233 amino acid paired homeodomain-
containing protein. The Ames dwarf mice had a missense mutation in the
homeodomain, p.S83P, that abrogates DNA binding. A null allele of Prop1 was
genetically engineered by inserting a LacZ/Neo cassette into the first exon of Prop1
(Nasonkin et al. 2004). These Prop1/ mice essentially phenocopy the Prop1df/df
mice when compared to the same genetic background. Prop1 expression is highly
restricted to the pituitary, beginning at embryonic day 10.5, reaching a peak at e12.5,
declining through development, and remaining detectable in the postnatal period
(Perez Millan et al. 2016; Sornson et al. 1996).
PROP1 binds β-catenin and competes with the HESX1-corepressor complex for
binding at the early enhancer of Pou1f1 to induce Pou1f1 expression (Dasen et al.
2001; Olson et al. 2006; Sornson et al. 1996). Deletion of β-catenin before the
expression of Pou1f1 blocks Pou1f1 expression, while deletion of β-catenin after the
onset of Pou1f1 expression had no effect, suggesting a narrow critical window of
β-catenin expression in lineage determination. This β-catenin phenotype occurs
independently of LEF1, TCF3, or TCF4 action. PROP1 interacts directly with the
corepressors TLE, Reptin, and HDAC1 to bind and repress the expression of Hesx1.
Lineage tracing reveals that all hormone-producing cell types of the anterior and
intermediate lobes of the pituitary descend from Prop1-expressing cells (Davis et al.
2016). This places Prop1 just downstream of Sox2 expression in pituitary
progenitors (Perez Millan et al. 2016) (Fig. 6.2). Prop1df/df mice have a vasculariza-
tion defect, reduced expression of cyclin D1 and cyclin D2 in pituitary progenitors,
and a failure of proliferating cells to migrate away from the stem cell niche into the
parenchyma, which appears to be a failed epithelial to mesenchymal-like transition
(Ward et al. 2005). Many mutant cells undergo apoptosis. Notch signaling is
normally active in the transitional zone between the stem cell niche and the
differentiating cells, but Prop1 mutants do not express Notch2 (Himes and Raetzman
2009; Raetzman et al. 2004). Colony forming assays and RNA-Seq were used to
assess the effect of Prop1 deficiency on stem cells. Mutants had a reduced stem cell
pool, abnormal colony morphology, and changes in gene expression consistent with
a failed epithelial-to-mesenchymal transition. To identify additional downstream
targets of Prop1 that might underlie the features of mutant pituitaries, Pérez Millán
and colleagues developed a biotin-tagged allele of PROP1 in an immortalized
pituitary cell line, facilitating direct identification of PROP1-binding sites in chro-
matin (Perez Millan et al. 2016). Binding sites were identified in or near Notch2,
Gli2, and Zeb2. Zeb2 activation appeared to be an essential step in the process of
engaging progenitors to undergo EMT.
6 Pituitary Development and Organogenesis: Transcription Factors in. . . 145
Quickly after the discovery of mouse Prop1 by positional cloning, the human
ortholog was identified and recessive loss of function mutations was found in four
unrelated families with CPHD (Wu et al. 1998). All three mutations either reduced or
completely eliminated PROP1’s ability to bind DNA, convincingly implicating
these lesions as the cause of CPHD. Mutations in PROP1 are the most common
known cause of CPHD (Cogan et al. 1998; Deladoey et al. 1999). The most frequent
mutations are frameshifts causing premature termination; many patients are homo-
zygous for c.[301-302delAG] or c.[150delA]. Portuguese (Lemos et al. 2006),
Lithuanian (Navardauskaite et al. 2014), and Czech (Lebl et al. 2005) cohorts have
a high incidence of the 2-bp deletion, and the 1-bp deletion is common in Croatia
(Dusatkova et al. 2016). A thorough investigation of the haplotype revealed that the
2-bp deletion founder mutation arose approximately 101 generations ago in central
eastern Europe, and it either reoccurred or was transferred to the Iberian peninsula
approximately 23 generations ago (Dusatkova et al. 2016). The Iberian haplotype
and mutation were subsequently transferred to Latin America. Combining all popu-
lation groups analyzed so far, PROP1 mutations are estimated to cause ~11% of
familial and sporadic cases of CPHD (Fang et al. 2016). All mutations discovered to
date are recessive, and the phenotype includes deficiency of GH, TSH,
gonadotropins, and progressive loss of ACTH. We speculate that the progressive
hormone loss may result from depletion of the pituitary stem cell pool.
Gh and Prl genes that were necessary for their expression in pituitary cell cultures
and used them as probes to screen a pituitary cDNA expression library (Ingraham
et al. 1988; Nelson et al. 1986). They identified the Pou1f1 cDNA, detected
transcripts only in the pituitary gland, and showed that Pou1f1 expression vectors
were sufficient to activate Gh and Prl reporter genes in transfected, heterologous
cells. Karin’s group took the approach of biochemically purifying the protein that
bound the cis-acting sequences in the Gh promoter and obtaining the amino acid
sequence of a peptide from the binding protein (Bodner et al. 1988; Bodner and
Karin 1987). Probes complementary to the peptide sequence were used to screen
pituitary cDNA libraries and GHF1 was identified. Loss of GHF1 resulted in the loss
of Gh expression (McCormick et al. 1988). Subsequent structure–function analyses
revealed the presence of the nuclear localization signal and the roles of the
homeodomain and POU-specific domain in specificity of DNA binding (Theill
et al. 1989).
Two years after linking the mouse dwarfism phenotype to lesions in Pou1f1, Khono
and colleagues identified the first POU1F1 mutation found in a patient with TSH,
GH, and prolactin deficiencies (Tatsumi et al. 1992). This c.C638T nonsense
mutation resulted in a truncation, p.R172X, N-terminal to the homeodomain of
POU1F1. Numerous POU1F1 mutations have been described in patients with
CPHD (reviewed in Fang et al. 2016). The majority are recessive loss of function
6 Pituitary Development and Organogenesis: Transcription Factors in. . . 147
mutations that present with deficiency of GH, PRL, and TSH. However, there are a
few notable exceptions. The p.R271W mutation is dominant negative and incom-
pletely penetrant, possibly due to variability in mono-allelic expression of the mutant
versus normal allele (Ohta et al. 1992; Okamoto et al. 1994; Radovick et al. 1992).
The p.R271W mutation interferes with binding to C/EBPα, and it permits recruit-
ment to centromeric heterochromatin (Enwright et al. 2003). The mutation also
interferes with tethering of POU1F1 dimers to the nuclear matrix via interactions
with β-catenin, SATB1, and matrin-3 (Skowronska-Krawczyk et al. 2014). A p.
K216E mutation in the homeodomain is also a dominant cause of CPHD (Cohen
et al. 1999). This mutation acts by interfering with the ability of POU1F1 to
autoactivate its late enhancer in response to retinoic acid stimulation (see below).
A p.P76L mutation in the transactivation domain causes dominant IGHD in humans
and recessive growth insufficiency in mice (Sobrier et al. 2016). TSH and PRL were
not measured in the mouse model. The amino acid substitution increases DNA
binding at the human GH1 locus control region and increases binding to several
transcription factors, including PITX1, LHX3a, and ELK1. The p.P76L change
decreases transactivation of Gh, but it has no effect on the Prl promoter. Two
other proline substitutions in the transactivation domain have been reported in
children with GH, PRL, and TSH deficiency: p.P14L and p.P24L (Fofanova et al.
1998; Ohta et al. 1992). Detailed functional studies were not carried out. The p.P14L
variant appears dominant with incomplete penetrance, and no inheritance data were
presented for the p.P24L allele. In sum, mutations in POU1F1 can cause CPHD or
IGHD and be recessive or dominant, and dominant mutations are informative about
the normal mechanism of action.
Pou1f1 expression is selectively high in the anterior lobe of the pituitary. The
cis-acting sequences necessary for Pou1f1 expression were mapped extensively in
transgenic mice (Rhodes et al. 1993). A 390-bp enhancer located ~10.4–10.2 kb
upstream of the Pou1f1 transcription start site is sufficient for Pou1f1 expression
postnatally in the context of the minimal 327 bp Pou1f1 promoter (DiMattia et al.
1997). This distal or “late enhancer” is bound by POU1F1 in an autoregulatory loop
necessary for maintaining expression. Another enhancer element located between
~8 kb upstream of Pou1f1 is bound by PROP1 and β-catenin, inducing initial
expression of Pou1f1, but this proximal or “early enhancer” is not sufficient for
the sustained expression of Pou1f1 into adulthood (DiMattia et al. 1997; Olson et al.
2006; Sornson et al. 1996).
The role of POU1F1 in establishing thyrotropes and Tshb expression was discovered
later (Charles et al. 2006; Dasen et al. 1999; Gordon et al. 1997). Ridgway and
148 A. Z. Daly and S. A. Camper
Mellon and colleagues identified a DNA element that regulates gonadotrope expres-
sion of Cga, which encodes the alpha-subunit of the heterodimeric glycoprotein
hormones, luteinizing hormone (LH), follicle-stimulating hormone (FSH), and
thyroid-stimulating hormone (TSH) (Horn et al. 1992). Mellon showed that this
element (which they called GSE for Gonadotrope-Specific Element) was bound by a
54-kD protein, which they labeled GSEB-1 (GSE-Binding factor 1). They
demonstrated that GSEB-1 is NR5A1 by showing that NR5A1 binds specifically
to the GSE in vitro and that NR5A1-specific antibodies antagonize GSEB-1 function
(Barnhart and Mellon 1994). This was the first evidence that NR5A1 had a role in the
pituitary gland.
Mellon discovered that NR5A1 regulates the transcription of Lhb. This
represented the second known direct target of NR5A1 in the pituitary gland.
NR5A1 binds a GSE in the promoter region of Lhb and increases its expression
(Halvorson et al. 1996; Horn et al. 1992). Nilson and colleagues showed that 776 bp
of the bovine LHB proximal promoter region was sufficient to drive gonadotrope-
specific expression in the transgenic mice (Keri et al. 1994). This sequence contained
the GSE, was responsive to GnRH regulation but was not androgen or estrogen
responsive. A mutation in the GSE element reduced transgene expression ten-fold
(Keri and Nilson 1996). This suggested that NR5A1 stimulates LHB expression by
binding the promoter-proximal GSE in cell culture and in vivo. Morohashi and
colleagues found two regulatory elements of Nr5a1, showing that NR5A1 binds one
of them to stimulate its own expression in adrenal cells (Honda et al. 1993), while
Clay and colleagues demonstrated that NR5A1 directly regulates expression of
Gnrhr via a GSE in the promoter-proximal region (Duval et al. 1997).
NR5A1 interacts with other transcription factors to promote gene expression. The
Drouin lab demonstrated that PITX1 and NR5A1 interact directly and synergize to
activate Lhb promoter but not Cga gene expression (Tremblay et al. 1998, 1999).
The PITX1 binding site is not required for this activity. NR5A1 also interacts with
the zinc-finger transcription factor EGR1 to regulate the expression of Lhb (Lee et al.
1996; Tremblay and Drouin 1999). Both NR5A1 and EGR1 binding sites are
required for the synergism. Milbrandt and Charnay independently demonstrated
6 Pituitary Development and Organogenesis: Transcription Factors in. . . 151
that EGR1 is necessary for Lhb expression in mice (Lee et al. 1996; Topilko et al.
1998). There are no exact obvious NR5A1 binding sites in the Fshb promoter, but
Mellon and colleagues found two sites bound by NR5A1 (Jacobs et al. 2003).
NR5A1 interacts with Nuclear Factor Y (NFYA) on the Fshb promoter to synergis-
tically activate Fshb transcription.
The first patient with an NR5A1 mutation was reported in 1999 (Achermann et al.
1999). The patient presented with primary adrenal failure, complete XY sex reversal,
and streak-like gonads. Her pituitary gland responded to GnRH stimulation. She was
heterozygous for a 2-bp mutation that causes a p.G35A missense in the first zinc
finger of NR5A1 that abrogates DNA binding. Her phenotype is consistent with
NR5A1’s role in adrenal function and stimulation of LH and FSH expression. The
sex reversal was surprising because mice heterozygous for a null allele of Nr5a1 are
phenotypically normal. Loss-of-function mutations have been associated with other
cases of 46XY sex reversal (Achermann et al. 2002; Correa et al. 2004; Lin et al.
2007; Mallet et al. 2004), 46XX sex reversal (Hasegawa et al. 2004; Swartz et al.
2017b), undervirilization (Swartz et al. 2017a), adrenocortical insufficiency (Biason-
Lauber and Schoenle 2000; Guran et al. 2016), premature ovarian failure (Lourenco
et al. 2009), and spermatogenic failure (Bashamboo et al. 2010). Most mutations
were dominant, some were associated with incomplete penetrance, and some were
recessive (Baetens et al. 2017; Bashamboo et al. 2016; Igarashi et al. 2017; Sekido
and Lovell-Badge 2008).
Mice homozygous for a hypomorphic Pitx2neo allele have no Nr5a1 expression and,
consequently, no Lhb, Fshb, or Gnrhr expression (Suh et al. 2002). This suggests
that Pitx2 is upstream of Nr5a1 in the transcriptional hierarchy regulating pituitary
development. Later, PITX2 was shown to bind a pituitary gonadotrope-specific
enhancer within intron 6 of the Nr5a1 gene (Shima et al. 2008). Several groups
used transgenic mice to localize elements necessary for Nr5a1 expression in various
tissues (Karpova et al. 2005; Shima et al. 2005; Stallings et al. 2002). The action of
Nr5a1 in promoting gonadotrope fate is antagonized by TBX19 (Pulichino et al.
2003b).
Pomc expression (Poulin et al. 1997). The detection of Neurod1 expression in the
developing pituitary gland from e12 to e16 is consistent with a role in initiating
Pomc expression. Neurod1 is transiently expressed in corticotropes, suggesting that
there are other factors involved in sustained Pomc transcription (Poulin et al. 2000).
PITX1 was identified in an effort to identify a trans-acting factor that bound
another cis-acting sequence within the promoter-proximal region of Pomc
(Lamonerie et al. 1996). This element was sufficient to confer reporter gene expres-
sion in AtT-20 cells but not heterologous cells. Drouin and colleagues used this
sequence to screen a cDNA expression library for DNA binding proteins, which led
to the discovery of a novel homeodomain transcription factor PTX, now known as
Pitx1. Pitx1 is expressed in the pituitary primordium, and is detected in all pituitary
cell types, but it is enriched in adult gonadotropes and thyrotropes (Charles et al.
2005; Lanctot et al. 1999). Mice homozygous for Pitx1 knockout die at birth and
have severe defects in mandible and hind limb development, but there is little effect
on pituitary development due to functional overlap with the related Pitx2 gene
(Charles et al. 2005; Lanctot et al. 1999; Szeto et al. 1999).
A cis-acting sequence adjacent to the PITX1 binding site is required for Pomc
expression, and the sequence motif, TCACACCA, is a T box transcription factor-
binding site (Lamolet et al. 2001). PCR amplification of Tbox cDNAs in AtT-20 cells
identified a novel factor, TPIT (officially known as TBX19), which is expressed early
in pituitary development and is enriched in corticotropes and melanotropes. Ectopic
expression of Tpit under the control of the Cga promoter in mice is sufficient to drive
Pomc expression, but not Neurod1 expression. Patients with isolated ACTH defi-
ciency were screened and two unrelated individuals were identified with mutations in
TPIT, confirming the critical role of this gene in POMC expression. Homozygous
deletion of Tpit results in almost complete loss of POMC in both the anterior and the
intermediate lobes, and the intermediate lobe is hypoplastic (Pulichino et al. 2003b).
Neurod1 expression is not obviously affected, suggesting that the commitment to
corticotrope and melanotrope fate has been initiated. In the absence of TPIT, the
progenitors in the intermediate lobe express Nr5a1 and the alpha and beta subunits of
TSH, FSH, and LH, suggesting that TPIT has a repressive effect on gonadotrope and
thyrotrope cell fates. TPIT interacts directly with NR5A1 to inhibit its transcriptional
activity. It is noteworthy that Pou1f1, Gh and Prl are not ectopically expressed in the
intermediate lobes of Tpit mutants. Overexpressing Tpit under the control of the Cga
promoter had little effect on Tshb expression, but it is sufficient to suppress expres-
sion of Nr5a1, Lhb, and Fshb, consistent with the idea that TPIT drives corticotrope
and melanotrope fate while suppressing gonadotrope fate.
TPIT mutations are the most common, known cause of congenital, isolated ACTH
deficiency (Couture et al. 2012; Metherell et al. 2004; Pulichino et al. 2003a;
Vallette-Kasic et al. 2005). Patients with homozygous loss-of-function mutations
typically present with undetectable plasma ACTH and corticosterone, and hypoplas-
tic adrenal glands. Affected babies may have cholestatic jaundice and potentially
154 A. Z. Daly and S. A. Camper
fatal hypoglycemia. Tpit null mice have more yellow pigment than wild-type mice,
due to the lack of MSH, but no obvious pigment defects are present in human
patients.
6.9.3 Summary
Pomc transcription and the corticotrope and melanotrope cell fates are promoted by
several transcription factors including TPIT, NeuroD1, PITX1, and PITX2. TPIT is a
T box transcription factor that binds promoter-proximal region of POMC and
induces its expression in concert with PITX factors. It is selectively expressed in
corticotropes and melanotropes, activating the corticotrope and melanotrope lineage
while repressing the gonadotrope lineage by antagonizing NR5A1 function in a
DNA binding-independent manner. Loss of TPIT results in neonatal-onset isolated
ACTH deficiency, but not juvenile onset.
6.10 PAX7
Pax7/ mice were generated more than 20 years ago and revealed the importance of
Pax7 for muscle and brain development (Jostes et al. 1990; Kassar-Duchossoy et al.
2005; Mansouri and Gruss 1998; Oustanina et al. 2004; Relaix et al. 2005, 2006;
Seale et al. 2000). The homozygous mutants die 3 weeks after birth, but the pituitary
gland was not characterized (Mansouri et al. 1996). Karlstrom and colleagues
showed that the SHH-responsive gene Pax7 is expressed in the melanotropes of
the developing zebra fish pituitary gland (Guner et al. 2008). Spatial and temporal
gene expression profiling revealed that Pax7 is expressed in progenitors that give
rise to intermediate lobe melanotropes in mice, and it was hypothesized to be a
critical factor for intermediate lobe cell fate through direct regulation of Notch
signaling (Budry et al. 2012; Goldberg et al. 2011; Hosoyama et al. 2010).
Drouin and colleagues showed that Pax7/ mice lack intermediate lobe
melanotropes, and the deficiency of Pax7 permits ectopic differentiation of
progenitors into corticotropes (Budry et al. 2012). Moreover, ectopic expression of
Pax7 in the immortalized, corticotrope-like AtT-20 cells is sufficient to convert them
to a melanotrope identity associated with both reduced expression of corticotrope-
enriched transcripts such as the NeuroD1, glucocorticoid receptor, CRH receptor,
and vasopressin receptor 1b and activation of melanotrope-enriched transcripts such
as the dopamine receptor Drd2 and prohormone convertase 2, Pcsk2, which cleaves
Pomc to produce MSH. Thus, Pax7 is a selector of intermediate lobe melanotrope
cell identity.
6 Pituitary Development and Organogenesis: Transcription Factors in. . . 155
6.10.4 Summary
PAX7 is the first validated pioneer transcription factor acting within the pituitary.
Important for both muscle and brain development, it was not implicated in pituitary
function until 2010. Melano/corticotrope precursor cells can become either cell type
in development, and the presence of PAX7 railroads the development to a
melanotrope fate whereas the absence of PAX7 results in a corticotrope fate.
PAX7 does this by making melanotrope-specific TPIT-binding sites available to
TPIT by binding to heterochromatin and converting it to euchromatin. The Drouin
lab has shown a very thorough mechanism by which PAX7 does this, revealing the
amount of time this chromatin-changing process takes, and showing its stability
through repeated DNA replication cycles. However, the exact complex required for
this change in the epigenetic landscape is unclear. Furthermore, all of the work done
to date has focused on the “opening” of DNA that is specific to melanotropes, while
it is yet unclear whether PAX7 “closes” corticotrope-specific DNA.
in adult pituitary. The steps involved in switching cells from their SOX2 expressing,
cycling state to their specific lineage is unclear. A thorough analysis of cell cycle
regulators from e11.5 to e16.5 helped clarify this, as it revealed three distinct
populations (Bilodeau et al. 2009). There is a cycling precursor population (defined
as Ki67-positive, and presumably SOX2-positive), a non-cycling precursor popula-
tion (defined as p57-positive and p27-negative), and a non-cycling differentiated
population (p57-negative and p27-positive). Drouin and colleagues showed that the
deletion of the major transcription factor TPIT resulted in an unusual population of
non-cycling, undifferentiated progenitors that were p57+ and p27+. This suggests
that p57 marks cell cycle exit, and major factors (PROP1, TPIT, NR5A1) silence
p57, permitting differentiation, and locking cells into their fate. Unlike the temporal
phasing of neuronal differentiation in the brain and retina, nearly all cells in the
pituitary leave the cell cycle at approximately the same time between e11.5 and
e13.5 (Davis et al. 2011). Two major questions remain: what triggers the transition
from the proliferating to non-proliferating state, and how are the major transcription
factors exclusively turned on in different progenitors? Notch, WNT, and YAP/TAZ
signaling play important roles in regulating progenitor proliferation (Cheung et al.
2018a; Gaston-Massuet et al. 2011; Lodge et al. 2019) The second question
concerning how each factor is turned on is almost certainly specific to the factor,
as common regulatory networks across cell lineages would fail to explain the highly
unique transcription network in the different cell types.
The prevailing model of cell-type differentiation in the pituitary involves an exit
from the cell cycle with a major transcription factor that inexorably pushes it down a
single path to induce a specific and exclusive fate. There is evidence to suggest,
however, that the differentiation is not as linear as once thought, and the fate may be
more plastic than the model suggests. Early analyses of the pituitary led Romeis to
propose a “one cell one hormone” model (von Möllendorff and Bargmann 1940).
However, a single gonadotrope can express and secrete both LH and FSH (Phifer
et al. 1973; Robyn et al. 1973) and GH and TSH are co-expressed in hypothyroid
mice (Horvath et al. 1990). Cells that express both GH and PRL may be a transient
population that represents a precursor of both somatotropes and lactotropes (Borrelli
et al. 1989; Frawley et al. 1985). Or, they may represent a transition from a
GH-expressing cell to PRL-expressing through the action of POU1F1, LSD1, and
ZEB1 (Wang et al. 2007). Boehm and colleagues discovered that at e17.5, nearly all
cells expressing FSH also express TSH, and that many of these cells express NR5A1
suggesting a close relationship between gonadotropes and thyrotropes (Wen et al.
2010). These studies highlight how closely related the various cell types are,
challenging the idea that major transcription factors direct an inflexible cell fate.
158 A. Z. Daly and S. A. Camper
Four major technological innovations are now available to gain insight into cell fate
decisions during development: innovative DNA library preparation methods cou-
pled with high-throughput sequencing, single-cell sequencing, CRISPR, and
organoids.
RNA sequencing of single cells (scRNA-seq) is now possible (Navin et al. 2011).
Various methods can be used to capture single cells including fluidics, manual
capture, and FACS (Shapiro et al. 2013), and there are different approaches for
obtaining the RNA sequences from individual cells, including bar coding RNA
library preparations from single cells followed by next-generation sequencing
(Zheng et al. 2017; Ziegenhain et al. 2017). scRNA-seq analysis algorithms allow
researchers to map out a shared differentiation pathway in an asynchronous popula-
tion of cells (Trapnell et al. 2014). Temporal changes in gene expression that occur
as differentiation proceeds can be inferred, revealing candidate genes that may drive
these differentiation pathways. The scRNA-seq methods are particularly valuable for
analysis of rare cell types or situations in which the relative contribution of individ-
ual cell types to the whole is shifting. Currently, very few publications containing
single-cell sequencing data from pituitary exist (Cheung et al. 2018b; Ruf-Zamojski
et al. 2018). But the work that has been done already shows that this approach is
powerful and can reveal previously unappreciated subpopulations of cells and novel
markers of specialized cell types.
6.12.3 CRISPR
Genetically engineered mice have been invaluable for understanding the role of
individual genes in pituitary development (Camper et al. 1995). Transgenic mice
160 A. Z. Daly and S. A. Camper
were used to identify the regulatory elements of the Gh gene (Jones et al. 1995) and
Nr5a1 (Karpova et al. 2005; Stallings et al. 2002) and to tease apart the differentia-
tion of the somatolactotrope (Borrelli et al. 1989). Homologous recombination in
embryonic stem cells was used to characterize the function of Hesx1 (Dasen et al.
2001) and numerous other pituitary genes, as described earlier. Despite the impor-
tance of these technologies, there are significant shortcomings, including expense
and inefficiency, and position effects associated with random integration of
transgenes. CRISPR-Cas9 (clustered regularly interspersed short palindromic
repeats-CRISPR-associated protein 9) technology has reduced the time and expense
needed to make genetic modifications in mammalian genomes (Cong et al. 2013).
Suddenly, knockouts, knockins, conditional alleles, and even more complex tissue-
specific CRISPR mice were developed. This technology will have a high impact on
the field as it is more broadly implemented.
6.12.4 Organoids
These powerful technological advances will help us clarify many questions that
remain in understanding pituitary development, namely understanding the switch
from proliferation to differentiation and the factors responsible for the differentiation
of the cell types. Single-cell sequencing will reveal more nuanced populations of
cells that will clarify the exact pathway from a Sox2-expressing progenitor cell,
through a non-cycling intermediate precursor, to a cell type that has some combina-
tion of major transcription factors and hormones, to finally, a differentiated cell type
that expresses high levels of hormones. Knowing exactly which factor is in which
cell will rigorously generate a model for differentiation that will be less prone to
revision. Highly enriched motifs and footprints in ATAC-seq data will implicate
transcription factors that are critical to this process, shortening the long list of factors
that will be generated by single-cell RNA-seq. ATAC-seq will also (in concert with
CHIP-seq for histone marks) find the regulatory elements for these major factors,
6 Pituitary Development and Organogenesis: Transcription Factors in. . . 161
Acknowledgments We thank Karine Rizzoti (Crick, London) for images of Sox2 mutant embryos,
Hironori Bando for critical reading of the manuscript, and the National Institutes of Health for
funding (T32GM007544 and T32HG000040 to AZD and HD034283 and HD097096 to SAC).
Key References
Sox2 deficiency impairs the differentiation of multiple pituitary cell types in humans
and mice (Kelberman et al. 2006). Sox2 expressing pituitary cells have the
potential to self-renew and generate all other hormone producing cells of the
pituitary gland (Fauquier et al. 2008). Sox2 expressing progenitors are recruited to
generate new hormone producing cells in adult animals challenged to increase
hormone production (Andoniadou et al. 2013).
Pitx1 was cloned as a transcription factor that binds a critical element of the Pomc
gene (Lamonerie et al. 1996). Pitx2 was cloned as a cause of Rieger syndrome
and in a search for homeobox genes expressed in the pituitary gland (Gage and
Camper 1997; Semina et al. 1996). Pitx1 deletion has a modest effect on pituitary
gonadotropes and thyrotropes (Szeto et al. 1999). Pitx2 has multiple roles in
pituitary development including early induction of Rathke’s pouch, differentia-
tion of gonadotropes and the Pou1f1 lineage, and overlapping function with Pitx1
(Charles et al. 2005; Gage et al. 1999).
Lhx3 knockout affects early pituitary development and causes apoptosis of the
pituitary primordium (Ellsworth et al. 2008; Sheng et al. 1996). Lhx4 knockout
affects early pituitary development and causes apoptosis in the pituitary
162 A. Z. Daly and S. A. Camper
primordium (Raetzman et al. 2002; Sheng et al. 1997). Lhx2 knockout affects
pituitary posterior lobe development (Zhao et al. 2010).
Hesx1 was identified as a transcription factor expressed in early development and
then concentrated in Rathke’s pouch (Hermesz et al. 1996). Hesx1 deficient mice
have multiple anomalies including pituitary hypoplasia, and severely affected
individuals have septo-optic dysplasia (Dattani et al. 1998). HESX1 suppresses
Prop1 expression through interactions with co-repressors, and in the presence of
beta-catenin, Prop1 activites Pou1f1 expression (Olson et al. 2006).
PROP1 is a pituitary-specific transcription factor that is necessary for activation of
Pou1f1 and for differentiation of thyrotropes, somatotropes and lactotropes
(Sornson et al. 1996). Prop1 is necessary for developing a normal pool of
pituitary stem cells, for driving an epithelial to mesenchymal-like transition
process through the activation of Zeb2 and suppression of cell adhesion (Perez
Millan et al. 2016). All hormone producing cells of the anterior and intermediate
lobes of the pituitary gland are derived from a Prop1 expressing progenitor
(Davis et al. 2016).
POU1F1 was cloned based on its ability to bind critical elements of the GH and PRL
genes (Bodner et al. 1988; Ingraham et al. 1988). Pou1f1 is mutated in mice that
lack somatotropes, lactotropes and thyrotropes (Camper et al. 1990; Li et al.
1990). POU1F1 interacts with GATA2 to drive thyrotrope fate, and Gata2 may
have overlapping function with Gata3 (Charles et al. 2006; Dasen et al. 1999).
Somatotropes and lactotropes are very similar cell types derived from POU1F1
expressing progenitors. Epigentic regulation involving LSD1 and ZEB1 affects
the specialization of progenitors to somatotropes and lactotropes (Wang et al.
2007). POU1F1 acts at the nuclear matrix to regulate gene expression
(Skowronska-Krawczyk et al. 2014).
NR5A1 is essential for the development of adrenal and multiple organs of the
reproductive axis including the pituitary gland gonadotropes (Ingraham et al.
1994; Zhao et al. 2010). NR5A1 interacts with GATA2 to promote gonadotrope
fate (Charles et al. 2006; Dasen et al. 1999).
TBX19 is critical for melanotrope and corticotrope fate and it antagonizes the action
of NR5A1 in promoting gonadotrope fate (Lamolet et al. 2001; Pulichino et al.
2003). Homozygous deletion of Tpit results in almost complete loss of POMC in
both the anterior and intermediate lobes, and the intermediate lobe is hypoplastic
(Pulichino et al. 2003).
Pax7/ mice lack intermediate lobe melanotropes, and the deficiency of Pax7
permits ectopic differentiation of progenitors into corticotropes (Budry et al.
2012). PAX7 is a pioneering transcription factor that opens chromatin permitting
access of transcription factors such as TPIT.
Single cell sequencing is a powerful technique for determining gene expression in
individual pituitary cell types without the need for genetic labeling or cell sorting
(Cheung et al. 2018b).
Pituitary organoids hold promise for studying pituitary development ex vivo
(Ozone et al. 2016; Suga et al. 2011).
6 Pituitary Development and Organogenesis: Transcription Factors in. . . 163
References
Achermann JC, Ito M, Ito M, Hindmarsh PC, Jameson JL (1999) A mutation in the gene encoding
steroidogenic factor-1 causes XY sex reversal and adrenal failure in humans. Nat Genet
22:125–126
Achermann JC, Ozisik G, Ito M, Orun UA, Harmanci K, Gurakan B, Jameson JL (2002) Gonadal
determination and adrenal development are regulated by the orphan nuclear receptor steroido-
genic factor-1, in a dose-dependent manner. J Clin Endocrinol Metab 87:1829–1833
Alvarado DM, McCall K, Aferol H, Silva MJ, Garbow JR, Spees WM, Patel T, Siegel M, Dobbs
MB, Gurnett CA (2011) Pitx1 haploinsufficiency causes clubfoot in humans and a clubfoot-like
phenotype in mice. Hum Mol Genet 20:3943–3952
Andersen B, Pearse RV 2nd, Jenne K, Sornson M, Lin SC, Bartke A, Rosenfeld MG (1995) The
Ames dwarf gene is required for Pit-1 gene activation. Dev Biol 172:495–503
Andoniadou CL, Matsushima D, Mousavy Gharavy SN, Signore M, Mackintosh AI, Schaeffer M,
Gaston-Massuet C, Mollard P, Jacques TS, Le Tissier P et al (2013) Sox2(+) stem/progenitor
cells in the adult mouse pituitary support organ homeostasis and have tumor-inducing potential.
Cell Stem Cell 13:433–445
Avilion AA, Nicolis SK, Pevny LH, Perez L, Vivian N, Lovell-Badge R (2003) Multipotent cell
lineages in early mouse development depend on SOX2 function. Genes Dev 17:126–140
Baek S, Sung MH (2016) Genome-scale analysis of cell-specific regulatory codes using nuclear
enzymes. Methods Mol Biol 1418:225–240
Baetens D, Stoop H, Peelman F, Todeschini AL, Rosseel T, Coppieters F, Veitia RA, Looijenga
LH, De Baere E, Cools M (2017) NR5A1 is a novel disease gene for 46,XX testicular and
ovotesticular disorders of sex development. Genet Med 19:367–376
Barnhart KM, Mellon PL (1994) The orphan nuclear receptor, steroidogenic factor-1, regulates the
glycoprotein hormone alpha-subunit gene in pituitary gonadotropes. Mol Endocrinol 8:878–885
Barr FG, Nauta LE, Davis RJ, Schafer BW, Nycum LM, Biegel JA (1996) In vivo amplification of
the PAX3-FKHR and PAX7-FKHR fusion genes in alveolar rhabdomyosarcoma. Hum Mol
Genet 5:15–21
Barski A, Cuddapah S, Cui K, Roh TY, Schones DE, Wang Z, Wei G, Chepelev I, Zhao K (2007)
High-resolution profiling of histone methylations in the human genome. Cell 129:823–837
Bartke A (1965) The response of two types of dwarf mice to growth hormone, thyrotropin, and
thyroxine. Gen Comp Endocrinol 5:418–426
Bashamboo A, Ferraz-de-Souza B, Lourenco D, Lin L, Sebire NJ, Montjean D, Bignon-Topalovic J,
Mandelbaum J, Siffroi JP, Christin-Maitre S et al (2010) Human male infertility associated with
mutations in NR5A1 encoding steroidogenic factor 1. Am J Hum Genet 87:505–512
Bashamboo A, Donohoue PA, Vilain E, Rojo S, Calvel P, Seneviratne SN, Buonocore F,
Barseghyan H, Bingham N, Rosenfeld JA et al (2016) A recurrent p.Arg92Trp variant in
steroidogenic factor-1 (NR5A1) can act as a molecular switch in human sex development.
Hum Mol Genet 25:5286
Bechtold-Dalla Pozza S, Hiedl S, Roeb J, Lohse P, Malik RE, Park S, Duran-Prado M, Rhodes SJ
(2012) A recessive mutation resulting in a disabling amino acid substitution (T194R) in the
LHX3 homeodomain causes combined pituitary hormone deficiency. Horm Res Paediatr
77:41–51
Bhangoo AP, Hunter CS, Savage JJ, Anhalt H, Pavlakis S, Walvoord EC, Ten S, Rhodes SJ (2006)
Clinical case seminar: a novel LHX3 mutation presenting as combined pituitary hormonal
deficiency. J Clin Endocrinol Metab 91:747–753
Biason-Lauber A, Schoenle EJ (2000) Apparently normal ovarian differentiation in a prepubertal
girl with transcriptionally inactive steroidogenic factor 1 (NR5A1/SF-1) and adrenocortical
insufficiency. Am J Hum Genet 67:1563–1568
Bilodeau S, Roussel-Gervais A, Drouin J (2009) Distinct developmental roles of cell cycle
inhibitors p57Kip2 and p27Kip1 distinguish pituitary progenitor cell cycle exit from cell
cycle reentry of differentiated cells. Mol Cell Biol 29:1895–1908
164 A. Z. Daly and S. A. Camper
Bodner M, Karin M (1987) A pituitary-specific trans-acting factor can stimulate transcription from
the growth hormone promoter in extracts of nonexpressing cells. Cell 50:267–275
Bodner M, Castrillo JL, Theill LE, Deerinck T, Ellisman M, Karin M (1988) The pituitary-specific
transcription factor GHF-1 is a homeobox-containing protein. Cell 55:505–518
Bonfig W, Krude H, Schmidt H (2011) A novel mutation of LHX3 is associated with combined
pituitary hormone deficiency including ACTH deficiency, sensorineural hearing loss, and short
neck-a case report and review of the literature. Eur J Pediatr 170:1017–1021
Borrelli E, Heyman RA, Arias C, Sawchenko PE, Evans RM (1989) Transgenic mice with inducible
dwarfism. Nature 339:538–541
Buckwalter MS, Katz RW, Camper SA (1991) Localization of the panhypopituitary dwarf mutation
(df) on mouse chromosome 11 in an intersubspecific backcross. Genomics 10:515–526
Budry L, Balsalobre A, Gauthier Y, Khetchoumian K, L’Honore A, Vallette S, Brue T, Figarella-
Branger D, Meij B, Drouin J (2012) The selector gene Pax7 dictates alternate pituitary cell fates
through its pioneer action on chromatin remodeling. Genes Dev 26:2299–2310
Buenrostro JD, Giresi PG, Zaba LC, Chang HY, Greenleaf WJ (2013) Transposition of native
chromatin for fast and sensitive epigenomic profiling of open chromatin, DNA-binding proteins
and nucleosome position. Nat Methods 10:1213–1218
Camper SA, Saunders TL, Katz RW, Reeves RH (1990) The Pit-1 transcription factor gene is a
candidate for the murine Snell dwarf mutation. Genomics 8:586–590
Camper SA, Saunders TL, Kendall SK, Keri RA, Seasholtz AF, Gordon DF, Birkmeier TS, Keegan
CE, Karolyi IJ, Roller ML et al (1995) Implementing transgenic and embryonic stem cell
technology to study gene expression, cell-cell interactions and gene function. Biol Reprod
52:246–257
Carreno G, Apps JR, Lodge EJ, Panousopoulos L, Haston S, Gonzalez-Meljem JM, Hahn H,
Andoniadou CL, Martinez-Barbera JP (2017) Hypothalamic sonic hedgehog is required for cell
specification and proliferation of LHX3/LHX4 pituitary embryonic precursors. Development
144:3289–3302
Carsner RL, Reynolds EG (1960) Primary site of gene action in anterior pituitary dwarf mice.
Science 131:829
Castinetti F, Reynaud R, Saveanu A, Jullien N, Quentien MH, Rochette C, Barlier A, Enjalbert A,
Brue T (2016) Mechanisms in endocrinology: an update in the genetic aetiologies of combined
pituitary hormone deficiency. Eur J Endocrinol 174(6):R239–R247
Cha KB, Douglas KR, Potok MA, Liang H, Jones SN, Camper SA (2004) WNT5A signaling affects
pituitary gland shape. Mech Dev 121:183–194
Chambers TJ, Giles A, Brabant G, Davis JR (2013) Wnt signalling in pituitary development and
tumorigenesis. Endocr Relat Cancer 20:R101–R111
Charles MA, Suh H, Hjalt TA, Drouin J, Camper SA, Gage PJ (2005) PITX genes are required for
cell survival and Lhx3 activation. Mol Endocrinol 19:1893–1903
Charles MA, Saunders TL, Wood WM, Owens K, Parlow AF, Camper SA, Ridgway EC, Gordon
DF (2006) Pituitary-specific Gata2 knockout: effects on gonadotrope and thyrotrope function.
Mol Endocrinol 20:1366–1377
Chen L, Gage PJ (2016) Heterozygous Pitx2 null mice accurately recapitulate the ocular features of
Axenfeld-Rieger syndrome and congenital glaucoma. Invest Ophthalmol Vis Sci 57:5023–5030
Chen J, Hersmus N, Van Duppen V, Caesens P, Denef C, Vankelecom H (2005) The adult pituitary
contains a cell population displaying stem/progenitor cell and early embryonic characteristics.
Endocrinology 146:3985–3998
Cheng TC, Beamer WG, Phillips JA 3rd, Bartke A, Mallonee RL, Dowling C (1983) Etiology of
growth hormone deficiency in little, Ames, and Snell dwarf mice. Endocrinology
113:1669–1678
Cheung L, Le Tissier P, Goldsmith SG, Treier M, Lovell-Badge R, Rizzoti K (2018a) NOTCH
activity differentially affects alternative cell fate acquisition and maintenance. Elife 7:e33318
6 Pituitary Development and Organogenesis: Transcription Factors in. . . 165
Cheung LYM, George AS, McGee SR, Daly AZ, Brinkmeier ML, Ellsworth BS, Camper SA
(2018b) Single-cell RNA sequencing reveals novel markers of male pituitary stem cells and
hormone-producing cell types. Endocrinology 159:3910–3924
Chou SJ, Hermesz E, Hatta T, Feltner D, El-Hodiri HM, Jamrich M, Mahon K (2006) Conserved
regulatory elements establish the dynamic expression of Rpx/HesxI in early vertebrate develop-
ment. Dev Biol 292:533–545
Cogan JD, Wu W, Phillips JA 3rd, Arnhold IJ, Agapito A, Fofanova OV, Osorio MG, Bircan I,
Moreno A, Mendonca BB (1998) The PROP1 2-base pair deletion is a common cause of
combined pituitary hormone deficiency. J Clin Endocrinol Metab 83:3346–3349
Cohen LE, Zanger K, Brue T, Wondisford FE, Radovick S (1999) Defective retinoic acid regulation
of the Pit-1 gene enhancer: a novel mechanism of combined pituitary hormone deficiency. Mol
Endocrinol 13:476–484
Cong L, Ran FA, Cox D, Lin S, Barretto R, Habib N, Hsu PD, Wu X, Jiang W, Marraffini LA et al
(2013) Multiplex genome engineering using CRISPR/Cas systems. Science 339:819–823
Correa RV, Domenice S, Bingham NC, Billerbeck AE, Rainey WE, Parker KL, Mendonca BB
(2004) A microdeletion in the ligand binding domain of human steroidogenic factor 1 causes
XY sex reversal without adrenal insufficiency. J Clin Endocrinol Metab 89:1767–1772
Couture C, Saveanu A, Barlier A, Carel JC, Fassnacht M, Fluck CE, Houang M, Maes M, Phan-
Hug F, Enjalbert A et al (2012) Phenotypic homogeneity and genotypic variability in a large
series of congenital isolated ACTH-deficiency patients with TPIT gene mutations. J Clin
Endocrinol Metab 97:E486–E495
Creyghton MP, Cheng AW, Welstead GG, Kooistra T, Carey BW, Steine EJ, Hanna J, Lodato MA,
Frampton GM, Sharp PA et al (2010) Histone H3K27ac separates active from poised enhancers
and predicts developmental state. Proc Natl Acad Sci USA 107:21931–21936
Cushman LJ, Burrows HL, Seasholtz AF, Lewandoski M, Muzyczka N, Camper SA (2000)
Cre-mediated recombination in the pituitary gland. Genesis 28:167–174
Dasen JS, O’Connell SM, Flynn SE, Treier M, Gleiberman AS, Szeto DP, Hooshmand F, Aggarwal
AK, Rosenfeld MG (1999) Reciprocal interactions of Pit1 and GATA2 mediate signaling
gradient-induced determination of pituitary cell types. Cell 97:587–598
Dasen JS, Martinez Barbera JP, Herman TS, Connell SO, Olson L, Ju B, Tollkuhn J, Baek SH, Rose
DW, Rosenfeld MG (2001) Temporal regulation of a paired-like homeodomain repressor/TLE
corepressor complex and a related activator is required for pituitary organogenesis. Genes Dev
15:3193–3207
Dateki S, Fukami M, Uematsu A, Kaji M, Iso M, Ono M, Mizota M, Yokoya S, Motomura K,
Kinoshita E et al (2010) Mutation and gene copy number analyses of six pituitary transcription
factor genes in 71 patients with combined pituitary hormone deficiency: identification of a single
patient with LHX4 deletion. J Clin Endocrinol Metab 95:4043–4047
Datson NA, Semina E, van Staalduinen AA, Dauwerse HG, Meershoek EJ, Heus JJ, Frants RR, den
Dunnen JT, Murray JC, van Ommen GJ (1996) Closing in on the Rieger syndrome gene on
4q25: mapping translocation breakpoints within a 50-kb region. Am J Hum Genet 59:1297–
1305
Dattani MT, Martinez-Barbera JP, Thomas PQ, Brickman JM, Gupta R, Martensson IL,
Toresson H, Fox M, Wales JK, Hindmarsh PC et al (1998) Mutations in the homeobox gene
HESX1/Hesx1 associated with septo-optic dysplasia in human and mouse. Nat Genet
19:125–133
Davis SW, Camper SA (2007) Noggin regulates Bmp4 activity during pituitary induction. Dev Biol
305:145–160
Davis SW, Mortensen AH, Camper SA (2011) Birthdating studies reshape models for pituitary
gland cell specification. Dev Biol 352:215–227
Davis SW, Keisler JL, Perez-Millan MI, Schade V, Camper SA (2016) All hormone-producing cell
types of the pituitary intermediate and anterior lobes derive from Prop1-expressing progenitors.
Endocrinology 157:1385–1396
Deladoey J, Fluck C, Buyukgebiz A, Kuhlmann BV, Eble A, Hindmarsh PC, Wu W, Mullis PE
(1999) “Hot spot” in the PROP1 gene responsible for combined pituitary hormone deficiency.
J Clin Endocrinol Metab 84:1645–1650
166 A. Z. Daly and S. A. Camper
DiMattia GE, Rhodes SJ, Krones A, Carriere C, O’Connell S, Kalla K, Arias C, Sawchenko P,
Rosenfeld MG (1997) The Pit-1 gene is regulated by distinct early and late pituitary-specific
enhancers. Dev Biol 182:180–190
Djakoure C, Guibourdenche J, Porquet D, Pagesy P, Peillon F, Li JY, Evain-Brion D (1996)
Vitamin A and retinoic acid stimulate within minutes cAMP release and growth hormone
secretion in human pituitary cells. J Clin Endocrinol Metab 81:3123–3126
Dusatkova P, Pfaffle R, Brown MR, Akulevich N, Arnhold IJ, Kalina MA, Kot K, Krzisnik C,
Lemos MC, Malikova J et al (2016) Genesis of two most prevalent PROP1 gene variants
causing combined pituitary hormone deficiency in 21 populations. Eur J Hum Genet
24:415–420
Duval DL, Nelson SE, Clay CM (1997) A binding site for steroidogenic factor-1 is part of a
complex enhancer that mediates expression of the murine gonadotropin-releasing hormone
receptor gene. Biol Reprod 56:160–168
Eicher EM, Beamer WG (1980) New mouse dw allele: genetic location and effects on lifespan and
growth hormone levels. J Hered 71:187–190
Ellsworth BS, Butts DL, Camper SA (2008) Mechanisms underlying pituitary hypoplasia and failed
cell specification in Lhx3-deficient mice. Dev Biol 313:118–129
Enwright JF 3rd, Kawecki-Crook MA, Voss TC, Schaufele F, Day RN (2003) A PIT-1
homeodomain mutant blocks the intranuclear recruitment of the CCAAT/enhancer binding
protein alpha required for prolactin gene transcription. Mol Endocrinol 17:209–222
Ericson J, Norlin S, Jessell TM, Edlund T (1998) Integrated FGF and BMP signaling controls the
progression of progenitor cell differentiation and the emergence of pattern in the embryonic
anterior pituitary. Development 125:1005–1015
Fang Q, George AS, Brinkmeier ML, Mortensen AH, Gergics P, Cheung LY, Daly AZ, Ajmal A,
Perez Millan MI, Ozel AB et al (2016) Genetics of combined pituitary hormone deficiency:
roadmap into the genome era. Endocr Rev 37:636–675
Fauquier T, Guerineau NC, McKinney RA, Bauer K, Mollard P (2001) Folliculostellate cell
network: a route for long-distance communication in the anterior pituitary. Proc Natl Acad
Sci USA 98:8891–8896
Fauquier T, Rizzoti K, Dattani M, Lovell-Badge R, Robinson IC (2008) SOX2-expressing progen-
itor cells generate all of the major cell types in the adult mouse pituitary gland. Proc Natl Acad
Sci USA 105:2907–2912
Fekete C, Lechan RM (2014) Central regulation of hypothalamic-pituitary-thyroid axis under
physiological and pathophysiological conditions. Endocr Rev 35:159–194
Filges I, Bischof-Renner A, Rothlisberger B, Potthoff C, Glanzmann R, Gunthard J, Schneider J,
Huber AR, Zumsteg U, Miny P et al (2012) Panhypopituitarism presenting as life-threatening
heart failure caused by an inherited microdeletion in 1q25 including LHX4. Pediatrics 129:
e529–e534
Fofanova OV, Takamura N, Kinoshita E, Yoshimoto M, Tsuji Y, Peterkova VA, Evgrafov OV,
Dedov II, Goncharov NP, Yamashita S (1998) Rarity of PIT1 involvement in children from
Russia with combined pituitary hormone deficiency. Am J Med Genet 77:360–365
Frawley LS, Boockfor FR, Hoeffler JP (1985) Identification by plaque assays of a pituitary cell type
that secretes both growth hormone and prolactin. Endocrinology 116:734–737
Gage PJ, Camper SA (1997) Pituitary homeobox 2, a novel member of the bicoid-related family of
homeobox genes, is a potential regulator of anterior structure formation. Hum Mol Genet
6:457–464
Gage PJ, Lossie AC, Scarlett LM, Lloyd RV, Camper SA (1995) Ames dwarf mice exhibit
somatotrope commitment but lack growth hormone-releasing factor response. Endocrinology
136:1161–1167
Gage PJ, Brinkmeier ML, Scarlett LM, Knapp LT, Camper SA, Mahon KA (1996) The Ames dwarf
gene, df, is required early in pituitary ontogeny for the extinction of Rpx transcription and
initiation of lineage-specific cell proliferation. Mol Endocrinol 10:1570–1581
6 Pituitary Development and Organogenesis: Transcription Factors in. . . 167
Gage PJ, Suh H, Camper SA (1999) Dosage requirement of Pitx2 for development of multiple
organs. Development 126:4643–4651
Gaston-Massuet C, Andoniadou CL, Signore M, Jayakody SA, Charolidi N, Kyeyune R, Vernay B,
Jacques TS, Taketo MM, Le Tissier P et al (2011) Increased wingless (Wnt) signaling in
pituitary progenitor/stem cells gives rise to pituitary tumors in mice and humans. Proc Natl
Acad Sci USA 108:11482–11487
Gergics P, Brinkmeier ML, Camper SA (2015) Lhx4 deficiency: increased cyclin-dependent kinase
inhibitor expression and pituitary hypoplasia. Mol Endocrinol 29:597–612
Giacomini D, Paez-Pereda M, Theodoropoulou M, Gerez J, Nagashima AC, Chervin A, Berner S,
Labeur M, Refojo D, Renner U et al (2006) Bone morphogenetic protein-4 control of pituitary
pathophysiology. Front Horm Res 35:22–31
Goldberg LB, Aujla PK, Raetzman LT (2011) Persistent expression of activated Notch inhibits
corticotrope and melanotrope differentiation and results in dysfunction of the HPA axis. Dev
Biol 358:23–32
Gordon DF, Lewis SR, Haugen BR, James RA, McDermott MT, Wood WM, Ridgway EC (1997)
Pit-1 and GATA-2 interact and functionally cooperate to activate the thyrotropin beta-subunit
promoter. J Biol Chem 272:24339–24347
Gordon DF, Tucker EA, Tundwal K, Hall H, Wood WM, Ridgway EC (2006) MED220/thyroid
receptor-associated protein 220 functions as a transcriptional coactivator with Pit-1 and GATA-
2 on the thyrotropin-beta promoter in thyrotropes. Mol Endocrinol 20:1073–1089
Gregory LC, Humayun KN, Turton JP, McCabe MJ, Rhodes SJ, Dattani MT (2015) Novel lethal
form of congenital hypopituitarism associated with the first recessive LHX4 mutation. J Clin
Endocrinol Metab 100:2158–2164
Gubbay J, Collignon J, Koopman P, Capel B, Economou A, Munsterberg A, Vivian N,
Goodfellow P, Lovell-Badge R (1990) A gene mapping to the sex-determining region of the
mouse Y chromosome is a member of a novel family of embryonically expressed genes. Nature
346:245–250
Guner B, Ozacar AT, Thomas JE, Karlstrom RO (2008) Graded hedgehog and fibroblast growth
factor signaling independently regulate pituitary cell fates and help establish the pars distalis and
pars intermedia of the zebrafish adenohypophysis. Endocrinology 149:4435–4451
Guran T, Buonocore F, Saka N, Ozbek MN, Aycan Z, Bereket A, Bas F, Darcan S, Bideci A, Guven
A et al (2016) Rare causes of primary adrenal insufficiency: genetic and clinical characterization
of a large Nationwide cohort. J Clin Endocrinol Metab 101:284–292
Gurnett CA, Alaee F, Kruse LM, Desruisseau DM, Hecht JT, Wise CA, Bowcock AM, Dobbs MB
(2008) Asymmetric lower-limb malformations in individuals with homeobox PITX1 gene
mutation. Am J Hum Genet 83:616–622
Halvorson LM, Kaiser UB, Chin WW (1996) Stimulation of luteinizing hormone beta gene
promoter activity by the orphan nuclear receptor, steroidogenic factor-1. J Biol Chem
271:6645–6650
Hasegawa T, Fukami M, Sato N, Katsumata N, Sasaki G, Fukutani K, Morohashi K, Ogata T
(2004) Testicular dysgenesis without adrenal insufficiency in a 46,XY patient with a heterozy-
gous inactive mutation of steroidogenic factor-1. J Clin Endocrinol Metab 89:5930–5935
Hermesz E, Mackem S, Mahon KA (1996) Rpx: a novel anterior-restricted homeobox gene
progressively activated in the prechordal plate, anterior neural plate and Rathke’s pouch of
the mouse embryo. Development 122:41–52
Higham CE, Johannsson G, Shalet SM (2016) Hypopituitarism. Lancet 388:2403–2415
Himes AD, Raetzman LT (2009) Premature differentiation and aberrant movement of pituitary cells
lacking both Hes1 and Prop1. Dev Biol 325:151–161
Ho Y, Tadevosyan A, Liebhaber SA, Cooke NE (2008) The juxtaposition of a promoter with a
locus control region transcriptional domain activates gene expression. EMBO Rep 9:891–898
Ho Y, Shewchuk BM, Liebhaber SA, Cooke NE (2013) Distinct chromatin configurations regulate
the initiation and the maintenance of hGH gene expression. Mol Cell Biol 33:1723–1734
168 A. Z. Daly and S. A. Camper
Karpova T, Presley J, Manimaran RR, Scherrer SP, Tejada L, Peterson KR, Heckert LL (2005) A
FTZ-F1-containing yeast artificial chromosome recapitulates expression of steroidogenic factor
1 in vivo. Mol Endocrinol 19:2549–2563
Kassar-Duchossoy L, Giacone E, Gayraud-Morel B, Jory A, Gomes D, Tajbakhsh S (2005) Pax3/
Pax7 mark a novel population of primitive myogenic cells during development. Genes Dev
19:1426–1431
Kelberman D, Rizzoti K, Avilion A, Bitner-Glindzicz M, Cianfarani S, Collins J, Chong WK, Kirk
JM, Achermann JC, Ross R et al (2006) Mutations within Sox2/SOX2 are associated with
abnormalities in the hypothalamo-pituitary-gonadal axis in mice and humans. J Clin Invest
116:2442–2455
Kelberman D, Rizzoti K, Lovell-Badge R, Robinson IC, Dattani MT (2009) Genetic regulation of
pituitary gland development in human and mouse. Endocr Rev 30:790–829
Keller-Wood M (2015) Hypothalamic-pituitary—adrenal axis-feedback control. Compr Physiol
5:1161–1182
Keri RA, Nilson JH (1996) A steroidogenic factor-1 binding site is required for activity of the
luteinizing hormone beta subunit promoter in gonadotropes of transgenic mice. J Biol Chem
271:10782–10785
Keri RA, Wolfe MW, Saunders TL, Anderson I, Kendall SK, Wagner T, Yeung J, Gorski J, Nett
TM, Camper SA et al (1994) The proximal promoter of the bovine luteinizing hormone beta-
subunit gene confers gonadotrope-specific expression and regulation by gonadotropin-releasing
hormone, testosterone, and 17 beta-estradiol in transgenic mice. Mol Endocrinol 8:1807–1816
Kioussi C, Briata P, Baek SH, Rose DW, Hamblet NS, Herman T, Ohgi KA, Lin C, Gleiberman A,
Wang J et al (2002) Identification of a Wnt/Dvl/beta-catenin!Pitx2 pathway mediating cell-
type-specific proliferation during development. Cell 111:673–685
Kitamura K, Miura H, Miyagawa-Tomita S, Yanazawa M, Katoh-Fukui Y, Suzuki R, Ohuchi H,
Suehiro A, Motegi Y, Nakahara Y et al (1999) Mouse Pitx2 deficiency leads to anomalies of the
ventral body wall, heart, extra- and periocular mesoderm and right pulmonary isomerism.
Development 126:5749–5758
Klopocki E, Kahler C, Foulds N, Shah H, Joseph B, Vogel H, Luttgen S, Bald R, Besoke R, Held K
et al (2012) Deletions in PITX1 cause a spectrum of lower-limb malformations including mirror-
image polydactyly. Eur J Hum Genet 20:705–708
Kragesteen BK, Brancati F, Digilio MC, Mundlos S, Spielmann M (2019) H2AFY promoter
deletion causes PITX1 endoactivation and Liebenberg syndrome. J Med Genet 56:246–251
Kristrom B, Zdunek AM, Rydh A, Jonsson H, Sehlin P, Escher SA (2009) A novel mutation in the
LIM homeobox 3 gene is responsible for combined pituitary hormone deficiency, hearing
impairment, and vertebral malformations. J Clin Endocrinol Metab 94:1154–1161
Kulig E, Camper SA, Kuecker S, Jin L, Lloyd RV (1998) Remodeling of hyperplastic pituitaries in
hypothyroid us-subunit knockout mice after thyroxine and 1713-estradiol treatment: role of
apoptosis. Endocr Pathol 9:261–274
Lafont C, Desarmenien MG, Cassou M, Molino F, Lecoq J, Hodson D, Lacampagne A,
Mennessier G, El Yandouzi T, Carmignac D et al (2010) Cellular in vivo imaging reveals
coordinated regulation of pituitary microcirculation and GH cell network function. Proc Natl
Acad Sci USA 107:4465–4470
Lala DS, Rice DA, Parker KL (1992) Steroidogenic factor I, a key regulator of steroidogenic
enzyme expression, is the mouse homolog of fushi tarazu-factor I. Mol Endocrinol 6:1249–1258
Lamolet B, Pulichino AM, Lamonerie T, Gauthier Y, Brue T, Enjalbert A, Drouin J (2001) A
pituitary cell-restricted T box factor, Tpit, activates POMC transcription in cooperation with Pitx
homeoproteins. Cell 104:849–859
Lamonerie T, Tremblay JJ, Lanctot C, Therrien M, Gauthier Y, Drouin J (1996) Ptx1, a bicoid-
related homeo box transcription factor involved in transcription of the pro-opiomelanocortin
gene. Genes Dev 10:1284–1295
Lanctot C, Gauthier Y, Drouin J (1999) Pituitary homeobox 1 (Ptx1) is differentially expressed
during pituitary development. Endocrinology 140:1416–1422
170 A. Z. Daly and S. A. Camper
Langlais D, Couture C, Balsalobre A, Drouin J (2012) The Stat3/GR interaction code: predictive
value of direct/indirect DNA recruitment for transcription outcome. Mol Cell 47:38–49
Lebl J, Vosahlo J, Pfaeffle RW, Stobbe H, Cerna J, Novotna D, Zapletalova J, Kalvachova B,
Hana V, Weiss V et al (2005) Auxological and endocrine phenotype in a population-based
cohort of patients with PROP1 gene defects. Eur J Endocrinol 153:389–396
Lee SL, Sadovsky Y, Swirnoff AH, Polish JA, Goda P, Gavrilina G, Milbrandt J (1996) Luteinizing
hormone deficiency and female infertility in mice lacking the transcription factor NGFI-A
(Egr-1). Science 273:1219–1221
Lemos MC, Gomes L, Bastos M, Leite V, Limbert E, Carvalho D, Bacelar C, Monteiro M,
Fonseca F, Agapito A et al (2006) PROP1 gene analysis in Portuguese patients with combined
pituitary hormone deficiency. Clin Endocrinol 65:479–485
Li S, Crenshaw EB 3rd, Rawson EJ, Simmons DM, Swanson LW, Rosenfeld MG (1990) Dwarf
locus mutants lacking three pituitary cell types result from mutations in the POU-domain gene
pit-1. Nature 347:528–533
Li H, Witte DP, Branford WW, Aronow BJ, Weinstein M, Kaur S, Wert S, Singh G, Schreiner CM,
Whitsett JA et al (1994) Gsh-4 encodes a LIM-type homeodomain, is expressed in the develop-
ing central nervous system and is required for early postnatal survival. EMBO J 13:2876–2885
Li M, Pevny L, Lovell-Badge R, Smith A (1998) Generation of purified neural precursors from
embryonic stem cells by lineage selection. Curr Biol 8:971–974
Lin CR, Kioussi C, O’Connell S, Briata P, Szeto D, Liu F, Izpisua-Belmonte JC, Rosenfeld MG
(1999) Pitx2 regulates lung asymmetry, cardiac positioning and pituitary and tooth morphogen-
esis. Nature 401:279–282
Lin L, Philibert P, Ferraz-de-Souza B, Kelberman D, Homfray T, Albanese A, Molini V, Sebire NJ,
Einaudi S, Conway GS et al (2007) Heterozygous missense mutations in steroidogenic factor
1 (SF1/Ad4BP, NR5A1) are associated with 46,XY disorders of sex development with normal
adrenal function. J Clin Endocrinol Metab 92:991–999
Lodge EJ, Santambrogio A, Russell JP, Xekouki P, Jacques TS, Johnson RL, Thavaraj S, Bornstein
SR, Andoniadou CL (2019) Homeostatic and tumourigenic activity of SOX2+ pituitary stem
cells is controlled by the LATS/YAP/TAZ cascade. Elife 8:e43996
Lourenco D, Brauner R, Lin L, De Perdigo A, Weryha G, Muresan M, Boudjenah R, Guerra-
Junior G, Maciel-Guerra AT, Achermann JC et al (2009) Mutations in NR5A1 associated with
ovarian insufficiency. N Engl J Med 360:1200–1210
Lu MF, Pressman C, Dyer R, Johnson RL, Martin JF (1999) Function of Rieger syndrome gene in
left-right asymmetry and craniofacial development. Nature 401:276–278
Luo X, Ikeda Y, Parker KL (1994) A cell-specific nuclear receptor is essential for adrenal and
gonadal development and sexual differentiation. Cell 77:481–490
Mallet D, Bretones P, Michel-Calemard L, Dijoud F, David M, Morel Y (2004) Gonadal dysgenesis
without adrenal insufficiency in a 46, XY patient heterozygous for the nonsense C16X mutation:
a case of SF1 haploinsufficiency. J Clin Endocrinol Metab 89:4829–4832
Mansouri A, Gruss P (1998) Pax3 and Pax7 are expressed in commissural neurons and restrict
ventral neuronal identity in the spinal cord. Mech Dev 78:171–178
Mansouri A, Stoykova A, Torres M, Gruss P (1996) Dysgenesis of cephalic neural crest derivatives
in Pax7-/- mutant mice. Development 122:831–838
Matsumoto A, Ishii S (1992) Atlas of endocrine organs: vertebrates and invertebrates. Springer,
Berlin
Mayran A, Khetchoumian K, Hariri F, Pastinen T, Gauthier Y, Balsalobre A, Drouin J (2018)
Pioneer factor Pax7 deploys a stable enhancer repertoire for specification of cell fate. Nat Genet
50:259–269
McCormick A, Wu D, Castrillo JL, Dana S, Strobl J, Thompson EB, Karin M (1988) Extinction of
growth hormone expression in somatic cell hybrids involves repression of the specific trans-
activator GHF-1. Cell 55:379–389
6 Pituitary Development and Organogenesis: Transcription Factors in. . . 171
Melamed P, Zhu Y, Tan SH, Xie M, Koh M (2006) Gonadotropin-releasing hormone activation of
c-jun, but not early growth response factor-1, stimulates transcription of a luteinizing hormone
beta-subunit gene. Endocrinology 147:3598–3605
Melmed S (2011) Pathogenesis of pituitary tumors. Nat Rev Endocrinol 7:257–266
Metherell LA, Savage MO, Dattani M, Walker J, Clayton PE, Farooqi IS, Clark AJ (2004) TPIT
mutations are associated with early-onset, but not late-onset isolated ACTH deficiency. Eur J
Endocrinol 151:463–465
Mollard P, Hodson DJ, Lafont C, Rizzoti K, Drouin J (2012) A tridimensional view of pituitary
development and function. Trends Endocrinol Metab 23:261–269
Morohashi K, Honda S, Inomata Y, Handa H, Omura T (1992) A common trans-acting factor,
Ad4-binding protein, to the promoters of steroidogenic P-450s. J Biol Chem 267:17913–17919
Mortensen AH, MacDonald JW, Ghosh D, Camper SA (2011) Candidate genes for
panhypopituitarism identified by gene expression profiling. Physiol Genomics 43:1105–1116
Nakano T, Ando S, Takata N, Kawada M, Muguruma K, Sekiguchi K, Saito K, Yonemura S,
Eiraku M, Sasai Y (2012) Self-formation of optic cups and storable stratified neural retina from
human ESCs. Cell Stem Cell 10:771–785
Nasonkin IO, Ward RD, Raetzman LT, Seasholtz AF, Saunders TL, Gillespie PJ, Camper SA
(2004) Pituitary hypoplasia and respiratory distress syndrome in Prop1 knockout mice. Hum
Mol Genet 13:2727–2735
Navardauskaite R, Dusatkova P, Obermannova B, Pfaeffle RW, Blum WF, Adukauskiene D,
Smetanina N, Cinek O, Verkauskiene R, Lebl J (2014) High prevalence of PROP1 defects in
Lithuania: phenotypic findings in an ethnically homogenous cohort of patients with multiple
pituitary hormone deficiency. J Clin Endocrinol Metab 99:299–306
Navin N, Kendall J, Troge J, Andrews P, Rodgers L, McIndoo J, Cook K, Stepansky A, Levy D,
Esposito D et al (2011) Tumour evolution inferred by single-cell sequencing. Nature 472:90–94
Nelson C, Crenshaw EB 3rd, Franco R, Lira SA, Albert VR, Evans RM, Rosenfeld MG (1986)
Discrete cis-active genomic sequences dictate the pituitary cell type-specific expression of rat
prolactin and growth hormone genes. Nature 322:557–562
Netchine I, Sobrier ML, Krude H, Schnabel D, Maghnie M, Marcos E, Duriez B, Cacheux V,
Moers A, Goossens M et al (2000) Mutations in LHX3 result in a new syndrome revealed by
combined pituitary hormone deficiency. Nat Genet 25:182–186
Ohta K, Nobukuni Y, Mitsubuchi H, Fujimoto S, Matsuo N, Inagaki H, Endo F, Matsuda I (1992)
Mutations in the Pit-1 gene in children with combined pituitary hormone deficiency. Biochem
Biophys Res Commun 189:851–855
Ohuchi H, Hori Y, Yamasaki M, Harada H, Sekine K, Kato S, Itoh N (2000) FGF10 acts as a major
ligand for FGF receptor 2 IIIb in mouse multi-organ development. Biochem Biophys Res
Commun 277:643–649
Okamoto N, Wada Y, Ida S, Koga R, Ozono K, Chiyo H, Hayashi A, Tatsumi K (1994) Monoallelic
expression of normal mRNA in the PIT1 mutation heterozygotes with normal phenotype and
biallelic expression in the abnormal phenotype. Hum Mol Genet 3:1565–1568
Olson LE, Tollkuhn J, Scafoglio C, Krones A, Zhang J, Ohgi KA, Wu W, Taketo MM, Kemler R,
Grosschedl R et al (2006) Homeodomain-mediated beta-catenin-dependent switching events
dictate cell-lineage determination. Cell 125:593–605
Oustanina S, Hause G, Braun T (2004) Pax7 directs postnatal renewal and propagation of myogenic
satellite cells but not their specification. EMBO J 23:3430–3439
Ozone C, Suga H, Eiraku M, Kadoshima T, Yonemura S, Takata N, Oiso Y, Tsuji T, Sasai Y (2016)
Functional anterior pituitary generated in self-organizing culture of human embryonic stem
cells. Nat Commun 7:10351
Peel MT, Ho Y, Liebhaber SA (2018) Transcriptome analyses of female Somatotropes and
Lactotropes reveal novel regulators of cell identity in the pituitary. Endocrinology
159:3965–3980
Perez Millan MI, Brinkmeier ML, Mortensen AH, Camper SA (2016) PROP1 triggers epithelial-
mesenchymal transition-like process in pituitary stem cells. Elife 5:e14470
172 A. Z. Daly and S. A. Camper
Perez C, Dastot-Le Moal F, Collot N, Legendre M, Abadie I, Bertrand AM, Amselem S, Sobrier
ML (2012) Screening of LHX2 in patients presenting growth retardation with posterior pituitary
and ocular abnormalities. Eur J Endocrinol 167:85–91
Pfaeffle RW, Savage JJ, Hunter CS, Palme C, Ahlmann M, Kumar P, Bellone J, Schoenau E,
Korsch E, Bramswig JH et al (2007) Four novel mutations of the LHX3 gene cause combined
pituitary hormone deficiencies with or without limited neck rotation. J Clin Endocrinol Metab
92:1909–1919
Pfaeffle RW, Hunter CS, Savage JJ, Duran-Prado M, Mullen RD, Neeb ZP, Eiholzer U, Hesse V,
Haddad NG, Stobbe HM et al (2008) Three novel missense mutations within the LHX4 gene are
associated with variable pituitary hormone deficiencies. J Clin Endocrinol Metab 93:1062–1071
Phifer RF, Midgley AR, Spicer SS (1973) Immunohistologic and histologic evidence that follicle-
stimulating hormone and luteinizing hormone are present in the same cell type in the human pars
distalis. J Clin Endocrinol Metab 36:125–141
Porter FD, Drago J, Xu Y, Cheema SS, Wassif C, Huang SP, Lee E, Grinberg A, Massalas JS,
Bodine D et al (1997) Lhx2, a LIM homeobox gene, is required for eye, forebrain, and definitive
erythrocyte development. Development 124:2935–2944
Poulin G, Turgeon B, Drouin J (1997) NeuroD1/beta2 contributes to cell-specific transcription of
the proopiomelanocortin gene. Mol Cell Biol 17:6673–6682
Poulin G, Lebel M, Chamberland M, Paradis FW, Drouin J (2000) Specific protein-protein
interaction between basic helix-loop-helix transcription factors and homeoproteins of the Pitx
family. Mol Cell Biol 20:4826–4837
Pow DV, Perry VH, Morris JF, Gordon S (1989) Microglia in the neurohypophysis associate with
and endocytose terminal portions of neurosecretory neurons. Neuroscience 33:567–578
Pulichino AM, Vallette-Kasic S, Couture C, Gauthier Y, Brue T, David M, Malpuech G, Deal C,
Van Vliet G, De Vroede M et al (2003a) Human and mouse TPIT gene mutations cause early
onset pituitary ACTH deficiency. Genes Dev 17:711–716
Pulichino AM, Vallette-Kasic S, Tsai JP, Couture C, Gauthier Y, Drouin J (2003b) Tpit determines
alternate fates during pituitary cell differentiation. Genes Dev 17:738–747
Qiao S, Nordstrom K, Muijs L, Gasparoni G, Tierling S, Krause E, Walter J, Boehm U (2016)
Molecular plasticity of male and female murine gonadotropes revealed by mRNA sequencing.
Endocrinology 157:1082–1093
Radovick S, Nations M, Du Y, Berg LA, Weintraub BD, Wondisford FE (1992) A mutation in the
POU-homeodomain of Pit-1 responsible for combined pituitary hormone deficiency. Science
257:1115–1118
Raetzman LT, Ward R, Camper SA (2002) Lhx4 and Prop1 are required for cell survival and
expansion of the pituitary primordia. Development 129:4229–4239
Raetzman LT, Ross SA, Cook S, Dunwoodie SL, Camper SA, Thomas PQ (2004) Developmental
regulation of Notch signaling genes in the embryonic pituitary: Prop1 deficiency affects Notch2
expression. Dev Biol 265:329–340
Ragge NK, Lorenz B, Schneider A, Bushby K, de Sanctis L, de Sanctis U, Salt A, Collin JR, Vivian
AJ, Free SL et al (2005) SOX2 anophthalmia syndrome. Am J Med Genet A 135:1–7.
discussion 8
Rajab A, Kelberman D, de Castro SC, Biebermann H, Shaikh H, Pearce K, Hall CM, Shaikh G,
Gerrelli D, Grueters A et al (2008) Novel mutations in LHX3 are associated with hypopituita-
rism and sensorineural hearing loss. Hum Mol Genet 17:2150–2159
Ramzan K, Bin-Abbas B, Al-Jomaa L, Allam R, Al-Owain M, Imtiaz F (2017) Two novel LHX3
mutations in patients with combined pituitary hormone deficiency including cervical rigidity
and sensorineural hearing loss. BMC Endocr Disord 17:17
Rappaport R, Amselem S (2001) Hypothalamic-pituitary development: genetic and clinical aspects.
Karger, Basel
Regal M, Paramo C, Sierra SM, Garcia-Mayor RV (2001) Prevalence and incidence of hypopitu-
itarism in an adult Caucasian population in northwestern Spain. Clin Endocrinol 55:735–740
6 Pituitary Development and Organogenesis: Transcription Factors in. . . 173
Zappone MV, Galli R, Catena R, Meani N, De Biasi S, Mattei E, Tiveron C, Vescovi AL, Lovell-
Badge R, Ottolenghi S et al (2000) Sox2 regulatory sequences direct expression of a (beta)-geo
transgene to telencephalic neural stem cells and precursors of the mouse embryo, revealing
regionalization of gene expression in CNS stem cells. Development 127:2367–2382
Zhao L, Bakke M, Krimkevich Y, Cushman LJ, Parlow AF, Camper SA, Parker KL (2001)
Steroidogenic factor 1 (SF1) is essential for pituitary gonadotrope function. Development
128:147–154
Zhao Y, Mailloux CM, Hermesz E, Palkovits M, Westphal H (2010) A role of the LIM-homeobox
gene Lhx2 in the regulation of pituitary development. Dev Biol 337:313–323
Zheng GX, Terry JM, Belgrader P, Ryvkin P, Bent ZW, Wilson R, Ziraldo SB, Wheeler TD,
McDermott GP, Zhu J et al (2017) Massively parallel digital transcriptional profiling of single
cells. Nat Commun 8:14049
Zhu X, Zhang J, Tollkuhn J, Ohsawa R, Bresnick EH, Guillemot F, Kageyama R, Rosenfeld MG
(2006) Sustained Notch signaling in progenitors is required for sequential emergence of distinct
cell lineages during organogenesis. Genes Dev 20:2739–2753
Zhu X, Gleiberman AS, Rosenfeld MG (2007) Molecular physiology of pituitary development:
signaling and transcriptional networks. Physiol Rev 87:933–963
Ziegenhain C, Vieth B, Parekh S, Reinius B, Guillaumet-Adkins A, Smets M, Leonhardt H,
Heyn H, Hellmann I, Enard W (2017) Comparative analysis of single-cell RNA sequencing
methods. Mol Cell 65(631–643):e634
Part II
Developmental Modulators and Epigenetic
Factors
Hypothalamic Development: Role of GABA
7
M. Stratton and S. Tobet
Abstract
A number of cellular developmental processes occur in an orchestrated fashion to
mediate the transition from a simple neural tube to the complexity of the adult
brain. A combination of internal programming and external cues provides the
molecular specificity needed to direct cells to the correct location, initiate correct
gene expression, and connect to the appropriate circuits. The heterogeneous cells
of the hypothalamus affect all aspects of physiology through the regulation of the
autonomic nervous system, motivated behavior, and endocrine balance. Small
disruptions (genetic or environmental) to hypothalamic development can impact
adult physiology and contribute to pathology. GABAergic signaling has drawn
the attention of developmental neuroendocrinologists as a potential effector
system of hypothalamic development. This chapter discusses the role of GABA
in directing the development of the hypothalamus, specifically related to the
migration of neurons that release gonadotropin-releasing hormone (GnRH), and
in the formation of the ventromedial (VMH) and paraventricular (PVN) nuclei of
the hypothalamus (Fig. 7.1).
Keywords
Neuron migration · Hypothalamic development · Gamma-aminobutyric acid
(GABA) · Gonadotropin-releasing hormone (GnRH) · Paraventricular nucleus of
the hypothalamus (PVN) · Ventromedial nucleus of the hypothalamus (VMH) ·
Developmental patterning
M. Stratton (*)
Department of Physiology and Cell Biology, Ohio State University, Columbus, OH, USA
e-mail: [email protected]
S. Tobet (*)
Department of Biomedical Sciences, School of Biomedical Engineering, Colorado State University,
Fort Collins, CO, USA
e-mail: [email protected]
Abbreviations
Fig. 7.1 Introduction to the role of GABA in hypothalamic development and potential
implications for adult physiology and pathophysiology. There is a striking absence of GABA and
GABA-producing enzymes inside select developing regions of the hypothalamus (e.g., VMH–
GAD67 depicted and PVN–GABA depicted) which is contrasted by enrichment of GABA
receptors within these regions (GABABR1 depicted). Subsequent text and figures summarize
work done to determine the role of GABA in directing the development of these regions and for
the GnRH system as well as potential consequences of altered GABA signaling during hypotha-
lamic development. The digital image in the background was taken from archival material in the
Tobet laboratory and shows immunoreactive GAD67 in adult mouse brain; dashed circles highlight
the VMH and PVN nuclei. Inset VMH images were adapted from Davis et al. (2002) and inset PVN
images were adapted from McClellan et al. (2010)
Fig. 7.2 Sagittal schematic diagram of an embryonic head illustrates hotspots for GABA regula-
tion of neuron migration in the developing hypothalamus. GABA has been shown to act generally
as a “brake” for hypothalamic cells as they exit nasal cavity and enter the brain (GnRH neurons), in
the developing ventromedial nucleus of the hypothalamus (VMH) and in the developing
paraventricular nucleus of the hypothalamus (PVN). GnRH neurons mostly track fibers of the
vomeronasal nerve from the vomeronasal organ (VNO) across the cortical plate (CP) toward the
accessory olfactory bulb (AOB). When a subset of fibers bifurcates and turn toward the basal
hypothalamus, GnRH neurons start making directional choices that are likely influenced by GABA
from neighboring GnRH neurons and from other neurons in the region. The other major source of
fibers is from main olfactory neuroepithelial cells (OE) that travel to the main olfactory bulb (OB).
For the PVN and VMH, cells originate medially and enter the region of the nuclei before they are
impacted by GABA at the boundaries of the nuclei as they coalesce
From a mechanistic perspective, there are two major organizational schemes in brain
development that individual cells must navigate; joining layers (e.g., cerebral cortex)
or groups (e.g., diencephalic nuclei). GnRH neurons indicate a less common third
case that does not do either (in most species) as these cells are not distributed in
either layer or confined to nuclear group. Due to the inability to categorize cells
solely based on distance from a proliferative zone, developmental studies of the
7 Hypothalamic Development: Role of GABA 185
hypothalamus may be intrinsically more complex than in areas like the cortex. As in
any tissue, the processes of proliferation, growth, migration, differentiation, and
death mediate the development of the hypothalamus. Process extension (axon and
dendrite) and circuit integration are additional crucial steps in brain development.
During development, each of these steps impacts the other and altered regulation of
one process can lead to profound effects in downstream processes. For instance, the
microenvironment and epigenetic program of the progenitor pool directs gene
expression which influences migration. The microenvironment through which a
cell migrates, further influences its epigenetic and gene expression program, which
influences its growth, differentiation, process extension, and circuit integration.
Regardless of whether you categorize functional entities of brain development
into larger units like prosomeres or split things into smaller units based on molecular
genetic characteristics (early chapters, Puelles and Rubenstein 2015), the business
end of development is operated by the molecules that provide signals from one cell
to another (Fig. 7.3). They can be recognition, adhesion or receptive entities on a cell
surface, or secreted, diffusive molecules that signal attraction or repulsion. The
current chapter addresses the secreted molecule GABA as a contributor to cell
movement and ultimately anatomical position. For secreted molecules to provide
patterning information they must have a way to address the issue of spatial distribu-
tion relative to nuclear boundaries that characterize the hypothalamus. Whereas a
radially migrating cell in the cerebral cortex can get 2-dimensional data to signal stop
or go relative to a layer, radially migrating cells in the hypothalamus must sense
curvilinear boundaries.
In regions outside of the hypothalamus, GABA had been suggested to elicit
multiple effects in neurodevelopment. These include effects on proliferation, migra-
tion, DNA synthesis, differentiation, growth, and morphology (reviewed in Wu and
Sun 2015). These developmental effects of GABA are thought to be mediated by
non-synaptic release/secretion of GABA and typically occur prior to the emergence
of structural synapses. Likely mechanisms for GABA release through non-synaptic
mechanisms include GABA transporters or through non-vesicular channel-mediated
release.
For hypothalamic cells, the first cells to leave the proliferative regions are usually the
ones that migrate the farthest. This is in contrast to structures such as the cerebral
cortex, where cells that populate the farthest layers, are the last ones to leave the
proliferative zone. Derived from the ventral and inferior lobes of the diencephalic
neuroepithelium, Altman and Bayer (1986) describe three waves of neurogenesis in
development of the rat hypothalamus. The first wave is between embryonic day (E)
13 and E15 (E0 ¼ first day post timed breeding; for mouse E11–E13) and these
neurons form lateral hypothalamic structures. The second wave is between E15 and
E17 with these neurons becoming part of the medial hypothalamus. The final wave
occurs between E17 and E19. These neurons make up the periventricular
186 M. Stratton and S. Tobet
Fig. 7.3 Schematic diagram illustrates a cellular perspective of the developmental boundaries of
the PVN and VMH that are defined by the presence/absence of GABA synthesizing cells. The cell
on the left in (a) suggests a cell found outside of the boundary that synthesizes and secretes GABA
following glutamic acid decarboxylase (GAD) acting on glutamate precursor. The cell on the right
in (a) suggests a cell found inside of the boundary that receives GABA through GABAA (forming a
chloride channel) or GABAB (G-protein coupled) receptors. Transcription factor programs initiated
by classic hypothalamic differentiation genes dictate cells that secrete and/or respond to GABA
signaling. Digital images in (b) depict the developmental boundaries of the VMH. Images demon-
strate GAD67 (left) and GABABR1 (right) immunoreactivities in the developing VMH, and were
adapted from Davis et al. (2002)
hypothalamic populations with notable exception of the PVN, for which most
neurons complete terminal mitosis prior to E15.
The field of birthdate labeling with tritiated thymidine to follow cells was
pioneered by investigators such as Sidman, Rakic, Altman, and Bayer. After 1988
(Miller and Nowakowski 1988), the field was gradually commandeered by studies
using bromodeoxyuridine (BrdU) because results could be generated more quickly
7 Hypothalamic Development: Role of GABA 187
using Nissl stains in mice until E17. However, many of the neurons that make up
these nuclear groups are present in groups earlier and express immunoreactive
peptides before E17. For example, immunoreactive corticotropin-releasing hormone
(CRH) and arginine vasopressin (AVP) can be found in the PVN by E15. It is not yet
known what event occurs between E15 and E17 to allow for Nissl stain detection,
although possible mechanisms include the proliferation of glia and/or tighter cell
packing mediated by migration. Since there is a GABAergic surround to both VMH
and PVN (see below), it is possible that GABA contributes to the fine-tuning of cell
positions that results in nuclear appearance.
The first requirement for connecting a signaling factor with hypothalamic develop-
ment is the demonstration of presence. Within the developing hypothalamus, multi-
ple immature cell groupings, including those destined to form PVN and VMH, are
enriched with GABA receptors while being bordered by areas with many cells
containing immunoreactive GABA or expressing GABA synthesizing enzymes
(Figs. 7.1 and 7.2). This striking pattern, first observed in the 1990s, provided the
first suggestions of an important role for GABA in directing hypothalamic
development.
GABA Sources: The patterning of GABA and its synthesizing enzymes can be
visualized efficiently through immunohistochemistry. Notably, GABA is found in
many cells throughout the hypothalamus where its synthetic enzymes glutamic acid
decarboxylases 1 and 2 (GAD aka GAD67 and GAD65) are found in cell bodies that
are excluded from locations within the developing VMH and PVN during develop-
ment (Tobet et al. 1999; McClellan et al. 2008; e.g., Fig. 7.1) and in adulthood (by in
situ hybridization for GAD; Okamura et al. 1990). Thus, the embryonic PVN and
VMH are ringed by regions of cells that contain immunoreactive GABA and GAD
while there is an absence of GABA and GAD within cells of the developing nuclear
groups. This pattern of expression also holds for GABA transporters GAT1 and
GAT4 (Jursky and Nelson 1996).
This void of GABAergic components in developing nuclei is apparent as early as
E12–13 in mice when the various developing nuclei are first detectable by
immunohistochemistry (but not yet Nissl). As GABAergic axons grow into these
hypothalamic nuclei, the pattern is lost (earlier for PVN than VMH) and the entire
hypothalamus becomes highly immunoreactive for GABAergic protein elements
soon after birth (GAD mRNA is still restricted to cells outside of the VMH and PVN;
Okamura et al. 1990). It is unclear if all of the sources of GABA are neuronal or if
some might be glial in nature. Thoughts of GAD as strictly neuronal might be only
7 Hypothalamic Development: Role of GABA 189
partially true as there is evidence of glial expression of GAD enzymes. Glia can also
synthesize GABA through the putrescine degradation pathway and is dependent on
the activity of monoamine oxidase B (Yoon and Lee 2014). More recent work also
points toward endothelial cells associated with developing blood vessels as being
sources of GABA in other brain regions (Li et al. 2018).
The case is less clear for patterning of GABAergic elements along the develop-
mental path of GnRH neurons. Consistent with this notion, a subset of GnRH
neurons (~30%) contain GABA during development (Tobet et al. 1996a) while
exposure to GABA varies across the long path of migration from the nasal compart-
ment to destinations in the hypothalamus and other brain regions. It is likely that
GABA from within and outside the population is one of the soluble secreted factors
that contribute to their migration.
of R1 and R2 subunits, for which multiple splice variants exist. GABAB receptor
activation has been shown to affect the adenylate cyclase system, decrease Ca2+
conductance, and increase K+ conductance. Originally thought to be mediated by
pertussis toxin-sensitive Giα and Goα G proteins, some pertussis toxin insensitive
baclofen (agonist) effects have been found in magnocellular neurons of PVN and
supraoptic nucleus (SON), in presynaptic receptors and whole-brain
co-immunoprecipitation mass spectrometry studies point toward Gβ1/β2, Gγ2, and
GαΖ as being predominant effector molecules (Schwenk et al. 2016).
The role of homeotic genes in the formation of the hypothalamus is complicated and
all the more so by a lack of understanding downstream molecular consequences.
When molecular genetics have been used to disrupt the expression of any of a
number of transcription factors, hypothalamic patterning has been dramatically
altered (see preceding chapters in this series). Whether the mechanisms are through
proliferation, migration, and terminal differentiation is often an open question.
Determining the impact on the molecules that mediate cell–cell communication
and induction of cellular processes is still less understood and infrequently
attempted. GABAergic signaling falls in this latter category.
Relative to understanding a role for GABA, a major homeodomain protein of
interest is the vertebrate homolog of the fly gene distal-less (DLX 1 and 2). DLX
protein exerts a direct regulatory impact on the transcription of glutamic acid
decarboxylases (GAD)65 and 67 (Le et al. 2017) and establishes the distribution
of GABA sources in multiple brain regions including the developing hypothalamus.
When a number of different transcription factors have been disrupted, a common
phenotype is aberrant Dlx expression and, presumably GABA. One homeobox-
containing gene, Orthopedia (Otp) is expressed in neurons that come to reside in
the PVN, SON, anterior periventricular, and arcuate nuclei. In mice in which Otp
was replaced by a lacZ reporter gene, cells migrated abnormally and were apparently
displaced by cells expressing Dlx1, in a manner suggesting that Dlx1-positive cells
replaced PVN-committed cells (positive for lacZ) that migrated abnormally
(Acampora et al. 1999). Interestingly, these types of knockout mice usually, when
they can be raised to adulthood, display altered physiology and behavior associated
with the hypothalamus including, response to dehydration, feeding, sexual function,
and stress responses. A key question is what downstream molecular changes follow
the transcription factor action(s) to alter cell positions? Are they secreted protein-
coding genes like sonic hedgehog or wnt’s or might they be GAD 65 or 67 leading to
changes in GABA that drive the movement of other cells?
The transcription factor HES1, a Notch effecter gene, is also expressed in the
developing pituitary and hypothalamus. In mice lacking HES1, distribution of GAD
67 immunoreactivity was shifted from surrounding the PVN to within the PVN. In
this case, AVP neurons were found to be located outside of the PVN boundaries in
the HES1 null (Aujla et al. 2011). The SON also showed ectopic GAD expression
7 Hypothalamic Development: Role of GABA 191
Fig. 7.4 Schematic diagram illustrates genetic manipulations that alter GABAergic cell
populations relative to boundaries in the developing ventromedial nucleus of the hypothalamus
(VMH) by different mechanisms. Loss of SF-1 function most likely causes cellular rearrangements
that result in GABAergic cells moving into the location of the VMH (Dellovade et al. 2000; Davis
et al. 2004). Loss of Rax function (knockout—KO) most likely causes cells in the location of the
VMH to change fate (transfate) to express glutamic acid decarboxylases (GAD) and make GABA
(Lu et al. 2013). In both cases, the phenotypes of cells in the VMH are altered and importantly,
GABA synthesizing cells become inappropriately located within the nucleus. Digital images from
sections taken from wild-type (WT) and SF1-KO mice were adapted from Dellovade et al. (2000).
The image of a section from a Rax KO mouse was adapted from one provided by Dr. D. Kurrasch
related to Lu et al. (2013)
in the HES1 null, which corresponded with the dispersion of typical AVP expressing
neurons in the SON. Based on the hypothesis that an inappropriately located cell is
exposed to alternate signaling milieus that have further functional consequences,
axons from these misplaced AVP neurons were unable to appropriately target the
posterior pituitary.
The gene Rax codes for a homeodomain protein that was found to be critical to
the formation of the VMH based on changes in the fates or molecular identities of
multiple cells (Lu et al. 2013). Mice lacking functional Rax also exhibited a change
in the pattern of GAD immunoreactivity. Cells containing GAD become localized
inside the VMH in contrast to its normal pattern of expression in cells surrounding
the nucleus. The authors pose two possible models to explain the Rax phenotype.
First, that mutant VMH cells “transfate” into a hypothalamic non-VMH identity, or
second that hypothalamic non-VMH cells (e.g., GAD containing) could migrate into
the mutant VMH region (Fig. 7.4). However, the former change in fate hypothesis
was considered more likely, based on the lack of change in the distribution of cells
expressing another transcription factor, Nkx2.1. Nonetheless, the new area of GAD
immunoreactivity within the VMH in the Rax null excludes several VMH cell types
that typically reside there.
When the transcription factor steroidogenic factor 1 (SF-1) was disrupted, the
positions of cells in and around the VMH change, including those expressing
192 M. Stratton and S. Tobet
Nkx2.1. Similar to Rax, cells that express GAD and normally surround the VMH
become localized within regions closer to the third ventricle and the distribution of
other markers of cell phenotype, including estrogen receptor-α, neuropeptide Y, and
galanin all become localized more medially (Dellovade et al. 2000). When the
amount of immunoreactive GAD was quantified across hypothalamic tissue
sections, the total amount of GAD did not appear to change. This suggests that
GAD expressing cells inappropriately invade the developing VMH in the SF-1 null,
rather than GAD expression being turned on inappropriately in typically VMH
resident cells. Also, cells that would normally express SF-1 and reside in the
VMH (assessed by SF-1-GFP transgenic reporter) become located more laterally
in the SF-1 null (Davis et al. 2004). The question in the SF-1 findings, like those
noted above, is identifying the mechanism behind the change. Transcription factors
are great markers for cell phenotypes and frequently code for genes that direct cell
differentiation, but to impact migration they must act by influencing genes that affect
cell movements. In this case, if GABA is an active agent influencing movement, and
if GABA acts as a stop signal for normal movement of SF-1 positive cells (more
discussion below), then when SF-1 is disrupted, GABA must no longer provide that
stop signal (Davis et al. 2004).
As alluded to above, a major tool for determining the impact of gene expression on
hypothalamic nuclear formation has been derived from animal models with selective
gene disruptions. It is particularly useful when the expression of a key regulatory
gene is somewhat selective for a given cell group. This is true for Sf-1 being
relatively restricted to the VMH and to some extent true for Sim-1, which is
somewhat restricted to the PVN (but does extend laterally and ventrally more so
than is usually discussed). Many of the other transcription factors discussed above
(e.g., Dlx) are expressed in many regions, but with patterns that still allow localized
interpretations when there are genetic disruptions. Unfortunately for GnRH neurons,
the only reliable marker is GnRH itself. Most of the genetic disruptions of transcrip-
tion factors that have been characterized as impacting GnRH neuron development,
do so indirectly, acting on elements other than the neurons themselves. However, an
advantage in this system is that most GnRH neurons found in immunohistochemical
studies can now be linked to one site of origin. Thus, genetic disruptions that
influence GnRH neuron movement can be characterized by changes in the distribu-
tion of GnRH immunoreactive cells at different developmental stages. The same
cannot be said of most other neurons in the brain. For example, neurons containing
CRH can arise from several different progenitor populations.
Much of the work done to investigate GABA’s role in directing neuron migration
in the hypothalamus has used a mouse line generated with YFP expressed under the
control of a modified neuronal Thy-1 promoter. In efforts to find a good reporter to
visualize synaptogenesis in motor neurons and retinal ganglion cells, 25 fluorescent
reporter lines were generated. Sixteen of these lines included YFP. Though the same
genetic element was used to drive fluorescent reporter expression, the lines have
7 Hypothalamic Development: Role of GABA 193
A key question posed in Sect. 7.3 is whether the molecular changes following
transcription factor actions involve secreted protein-coding genes like sonic hedge-
hog or wnt’s or might they be acting through GAD 65/67 leading to changes in
GABA that drive the movement of other cells? This is not a question easily answered
in vivo. To get a better look at the movement of cells in real-time and in response to
specific factors that influence motion, in vitro methods are needed. There are two
usual methods to evaluate influences on motion. In one case, dispersed cell cultures
have been used. This is perhaps most appropriate when there is a relatively homog-
enous population to work with, such as immortalized cell lines that represent GnRH
neurons (e.g., NLT, GT-1). Transwell migration assays and wound-healing migra-
tion assays were both able to demonstrate roles for secretory products impacting
NLT cells in vitro (reviewed Wierman et al. 2011). The problem for GnRH neurons
in vitro and for many other regions in the hypothalamus is the cellular diversity and
the dependence of cell behaviors on the surrounding cellular environment. For
GABA, the transwell migration assays have been done with cerebral cortical cells
(e.g., Behar et al. 1998), but were also complemented by later studies using
organotypic slice cultures (e.g., Behar et al. 2000).
Organotypic cultures have an advantage when cell diversity and local environ-
ment is important for understanding influences. It is important to note that in the
hypothalamus, the plane of sectioning is critical as the tissue slice should contain the
source of the migrating cells, the destination, and any cues used by the cell to direct
migration. These are different planes of sectioning for radial hypothalamic migration
(coronal) versus migration of GnRH neurons, which originate in the nasal compart-
ment (sagittal). Tissue explants offer relatively intact systems to test hypotheses of
cellular behavior but suffer limitations of inclusion of all necessary cues, imaging
hurdles for depth, as well as tissue health deep in tissue that has been removed from
nutrient and oxygen vascular perfusion. Two approaches have been used for creating
in vitro models of GnRH neuron migration. The first was an explant model (Fueshko
and Wray 1994) and then a head slice model (Tobet et al. 1996b). Both models have
been used to test the role of GABA; first in explants and then in head slices (see
Wierman et al. 2011). Live views of GABA influences on GnRH neurons were
discerned in slices (Bless et al. 2005) and then in explants (Casoni et al. 2012). The
slice model maintained the arrangements of axonal fiber guides as they existed
in vivo, while the explant model allows free extension of those axons away from
their nasal origins. Each technique has its place, and all have been successfully used
to shed light on many aspects of neurodevelopment.
194 M. Stratton and S. Tobet
Fig. 7.5 Schematic diagram illustrates GABAA receptor activation decreasing GnRH neuron exit
from the nasal compartment. (a) Montage shows immunochemically labeled GnRH neurons in
migration in a mouse sagittal brain section at E15 (adapted from Tobet et al. 2001). (b) GnRH
neurons originate in the vomeronasal organ (VNO) and migrate toward the brain along olfactory
axons, crossing the cribriform plate. (c) When slice cultures are treated with a GABAA receptor
agonist, fewer GnRH neurons cross the cribriform plate. Additional molecular complexity is present
in this system: GABA-producing GnRH neurons are shown in red and GABAA receptor containing
cells marked with red tags. BF Basal forebrain, OE olfactory epithelia, AOB accessory olfactory
bulb, OB olfactory bulb
7 Hypothalamic Development: Role of GABA 195
olfactory bulb and increased their number in the nasal compartment. Meanwhile,
bicuculline treatment resulted in GnRH neurons more scattered in the brain and
olfactory bulb (Bless et al. 2000). When fluorescence video microscopy was used to
measure acute effects of GABA signaling on GnRH cell movement, GABAA
receptor blockade resulted in increased migration and decreased turning (Bless
et al. 2005). These results were confirmed in explants where picrotoxin, a GABAA
receptor antagonist increased total distance traveled for GnRH neurons while
muscimol again restricted movement (Casoni et al. 2012). Interestingly, the work
addressing GABA’s role in GnRH cell movements was preceded by work showing
that GABA regulates GnRH electrophysiology in explant cultures started from
mouse embryos at E11.5 (Kusano et al. 1995) when GnRH neuron birth and
migration is just beginning. Interpretations of GABA effects on GnRH migration
are complicated by the fact that ~30% of migrating GnRH neurons are themselves
GABA positive and thus are not simply receivers of GABA cues (see genetic
manipulation below).
bicuculline, movement patterns in these areas switched to being more parallel to the
orientation of radial fibers. This data, in agreement with work done in the cortex
(Behar et al. 1998), indicates that the functional result of GABA signaling can
change based on location, GABA concentration, or developmental stage. GABA
signaling through the GABAB receptor also has a strong influence on VMH devel-
opment. Activation of GABAB receptors with baclofen decreased cell movements in
the VMH (Davis et al. 2002) while inhibition with saclofen increased cell movement
speed (McClellan et al. 2008).
VMH: Genetics Many GABAA receptor subunits are expressed in the developing
hypothalamus, some of which are expressed in subregions. As the β3 subunit is
highly expressed in the VMH and its disruption was known to reduce GABAA
receptor binding by 50% across the brain, β3 subunit-disrupted mice were a logical
place to start dissecting the role of GABAA receptors from a genetic perspective.
Mice lacking GABAA β3 subunits had larger VMH areas and cells containing ERα
were more widely dispersed than in controls (Dellovade et al. 2001). In mice lacking
functional GABAB receptors, the size of the VMH also was larger and there was
increased dorsal-to-ventral spread of cells containing ERα in the ventrolateral
quadrant (McClellan et al. 2008). Both of these findings are consistent with the
hypothesis that GABA produced at the lateral edges of the VMH provides a slowing
or stop signal (Fig. 7.4).
PVN: Genetics It was not surprising that the disruption of the β3 subunit of the
GABAA receptor did not impact the PVN as β3 subunit expression in the PVN was
found to be quite low (McClellan et al. 2010). On the other hand, in mice lacking
functional GABAB receptors, the cytoarchitecture of the PVN was altered. Specifi-
cally, relative to littermate controls, cells containing nNOS were less tightly packed,
cells containing ERα are found more laterally, and there was decreased expression of
BDNF in the PVN (McClellan et al. 2010) while CRH expression was increased
(Stratton et al. 2011). There was an unexpected finding of reduced blood vessel
density in these mice (Frahm et al. 2012) that provided a new target and new
7 Hypothalamic Development: Role of GABA 197
GABA Receptor Synergy? As alluded to above, GABA may influence cell posi-
tioning by affecting the guidance of cells by neuronal (as per GnRH; Bless et al.
2000) or glial (as per VMH; Dellovade et al. 2001) fibers. Manipulation of GABAA
receptor signaling altered the orientation of cell movements (Dellovade et al. 2001),
whereas GABAB receptor signaling decreased the rate of cell movement (Davis et al.
2002). Since the two receptors may be differentially sensitive to GABA levels, the
interaction of the two mechanisms may be spatially regulated by the locations of
different concentrations of GABA (e.g., high at the nuclear boundary). It is still
unknown whether GABAA and GABAB receptors act synergistically, antagonisti-
cally, or independently to influence movements in the developing hypothalamus.
7.8 Perspectives
particular functional endpoint at a later time point in the animal’s life. Although it
may be easier to tie a phenotype to a particular time point with chemical targeting, if
the drug is given systemically, it is still difficult to prove the functional phenotype is
brought about via chemical activity at that particular location of interest.
Despite these limitations, there is sufficient evidence to suspect phenotypes/
disorders that are related to hypothalamic function (and other brain regions) in
humans subjected to in utero exposure to drugs that manipulate GABA signaling.
Exposures to barbiturates, benzodiazepines, and other drugs (e.g., valproate) that
impact GABAergic systems can occur when the mother is being treated for epilepsy,
anxiety, or depression disorders. This is particularly true for pregnancies prior to the
mid-1980s when evidence for potential effects on the fetus began to mount. While
these drugs are at times, still used during pregnancy, it is after an informed risk–
benefit analysis. Most studies assessing effects in the fetus are limited in time of
follow-up and typically do not extend into the post-pubescent period (or even middle
age) when subtle alterations in hypothalamic function may manifest. In one rare
study that followed the offspring of mothers treated with barbiturates, potential
effects were seen in sexual preference and gender identity (Dessens et al. 1999).
What other phenotypes might be seen? For cells in hypothalamic cell groups such
as the VMH, PVN, and ARC that are all part of the classic feeding/obesity circuitry,
it is easy to consider that as a functional endpoint. The PVN also regulates systemic
glucocorticoids and autonomic tone in the cardiovascular system. Thus, develop-
mental GABA signaling in the hypothalamus represents a potential mechanism that
might link the highly comorbid conditions of depression, cardiovascular disease
(CVD), and obesity (Tobet et al. 2013). Comorbidity may point toward interactions
in adult physiology, for which obvious mechanisms exist between obesity and CVD,
and mechanisms linking depression to CVD or obesity are being developed
(Fig. 7.6). Nonetheless, a common developmental root to all three may lie in
common hypothalamic or even GABAergic etiology. Strikingly, Prader–Willi Syn-
drome (PWS) includes in its spectrum of symptoms, depression, obesity, CVD, and
reproductive phenotypes that could be due to effects in the GnRH system, and
additional behavioral phenotypes potentially rooted in the VMH. Dysregulation of
multiple hypothalamic circuits/functions is well documented in PWS patients.
Though the exact molecular roots of PWS are complex and still under investigation,
the spectrum of symptoms may well represent the potential phenotypes, albeit
severe, to be expected in humans subjected to conditions that impact the function
of GABAergic systems in the hypothalamus during early development.
(continued)
200 M. Stratton and S. Tobet
Fig. 7.6 Schematic diagram illustrates potential comorbid consequences of altered GABAergic
signaling. Hypothalamic dysfunction may provide a shared anatomical substrate for clinical
comorbidity. There are high rates of comorbidity among depression, cardiovascular disease,
obesity, and reproductive dysfunction. Altered hypothalamic development through genetic or
environmental insult could predispose individuals to one or more of the above conditions. Once
present, pathophysiology in one system is likely to disrupt physiology of the other systems and
cause additional disease. Brain circuits that regulate mood, cardiovascular function, metabolism,
and sexual function are all connected within the hypothalamus
7 Hypothalamic Development: Role of GABA 201
Fig. 7.7 The “GABA Shunt” uses α-Ketoglutarate, an intermediate of the TCA cycle, to produce
glutamic acid, which is decarboxylated by GAD enzymes to GABA. The 4-carbon chain of GABA
can re-enter the TCA cycle following conversion to succinic semialdehyde and ultimately succinate.
GABA is also produced in putrescine degradation or again from glutamic acid but derived from
glutamine. Many of these reactions are also reversible (e.g. glutamic acid can be converted to
α-ketoglutarate)
Acknowledgments We would like to thank Dr. Kristy McClellan for helpful discussion, Mr. Luke
Schwerdtfeger for expert assistance in preparing figures, and Dr. Deborah Kurrasch for providing an
original digital image for use in Fig. 7.4. MS was supported by 1K01AG056848.
Key References
Altman and Bayer (1986). This is a classic description of the birth of cells in the
diencephalon.
Frahm et al. (2012). First paper to indicate a role for regulated development of
hypothalamic vasculature.
Herbison et al. (2008). This paper is notable for being the first to directly address the
redundancy in the GnRH population needed for fertility.
Miller and Nowakowski (1988). This paper introduced the BrdU method to the
neuroscience community and kicked off a plethora of studies of cell proliferation
in the nervous system due to the relative ease compared to the use of tritiated
thymidine.
Okamura et al. (1990). This paper was the first to note the special relationship
between cell bodies for GABA synthesis based on the localization of glutamic
acid decarboxylases and how they lie outside of major hypothalamic cell groups.
202 M. Stratton and S. Tobet
Puelles and Rubenstein (2015). This paper updates history and approaches to
parcellation of the hypothalamus on the basis of large anatomical assumptions
to smaller units informed by molecular data.
Rakic (1972). This paper defined the classic view of radial glia providing migratory
guidance for neuroblasts in cerebral cortex.
Salic and Mitchison (2008). This paper provides an impetus to switch from BrdU to
EdU as a marker for DNA synthesis and thereby (indirectly) proliferation.
References
Acampora D, Postiglione MP, Avantaggiato V, Di Bonito M, Vaccarino FM, Michaud J, Simeone
A (1999) Progressive impairment of developing neuroendocrine cell lineages in the hypothala-
mus of mice lacking the Orthopedia gene. Genes Dev 13(21):2787–2800
Altman J, Bayer SA (1986) The development of the rat hypothalamus. Advances in anatomy,
embryology and cell biology. Springer, Berlin. Monograph
Aujla PK, Bora A, Monahan P, Sweedler JV, Raetzman LT (2011) The Notch effector gene Hes1
regulates migration of hypothalamic neurons, neuropeptide content and axon targeting to the
pituitary. Dev Biol 353(1):61–71
Behar TN, Schaffner AE, Scott CA, O’Connell C, Barker JL (1998) Differential response of cortical
plate and ventricular zone cells to GABA as a migration stimulus. J Neurosci 18(16):6378–6387
Behar TN, Schaffner AE, Scott CA, Greene CL, Barker JL (2000) GABA receptor antagonists
modulate postmitotic cell migration in slice cultures of embryonic rat cortex. Cereb Cortex
10(9):899–909
Bessman SP, Rosen J, Layne EC (1953) Gamma-aminobutyric acid-glutamic acid transamination in
brain. J Biol Chem 201(1):385–391
Bless EP, Westaway WA, Schwarting GA, Tobet SA (2000) Effects of gamma-aminobutyric acid
(A) receptor manipulation on migrating gonadotropin-releasing hormone neurons through the
entire migratory route in vivo and in vitro. Endocrinology 141(3):1254–1262
Bless EP, Walker HJ, Yu KW, Knoll JG, Moenter SM, Schwarting GA, Tobet SA (2005) Live view
of gonadotropin-releasing hormone containing neuron migration. Endocrinology
146(1):463–468
Casoni F, Hutchins BI, Donohue D, Fornaro M, Condie BG, Wray S (2012) SDF and GABA
interact to regulate axophilic migration of GnRH neurons. J Cell Sci 125(Pt 21):5015–5025
Davis AM, Henion TR, Tobet SA (2002) Gamma-aminobutyric acid B receptors and the develop-
ment of the ventromedial nucleus of the hypothalamus. J Comp Neurol 449(3):270–280
Davis AM, Seney ML, Stallings NR, Zhao L, Parker KL, Tobet SA (2004) Loss of steroidogenic
factor 1 alters cellular topography in the mouse ventromedial nucleus of the hypothalamus.
J Neurobiol 60(4):424–436
Dellovade TL, Young M, Ross EP, Henderson R, Caron K, Parker K, Tobet SA (2000) Disruption
of the gene encoding SF-1 alters the distribution of hypothalamic neuronal phenotypes. J Comp
Neurol 423(4):579–589
Dellovade TL, Davis AM, Ferguson C, Sieghart W, Homanics GE, Tobet SA (2001) GABA
influences the development of the ventromedial nucleus of the hypothalamus. J Neurobiol
49(4):264–276
Dessens AB, Cohen-Kettenis PT, Mellenbergh GJ, vd Poll N, Koppe JG, Boer K (1999) Prenatal
exposure to anticonvulsants and psychosexual development. Arch Sex Behav 28(1):31–44
Frahm KA, Schow MJ, Tobet SA (2012) The vasculature within the paraventricular nucleus of the
hypothalamus in mice varies as a function of development, subnuclear location, and GABA
signaling. Horm Metab Res 44(8):619–624
7 Hypothalamic Development: Role of GABA 203
Roberts E, Frankel S (1951) Glutamic acid decarboxylase in brain. J Biol Chem 188:789–795
Salic A, Mitchison TJ (2008) A chemical method for fast and sensitive detection of DNA synthesis
in vivo. Proc Natl Acad Sci U S A 105(7):2415–2420
Schwanzel-Fukuda M, Pfaff DW (1989) Origin of luteinizing hormone-releasing hormone neurons.
Nature 338(6211):161–164
Schwenk J, Pérez-Garci E, Schneider A, Kollewe A, Gauthier-Kemper A, Fritzius T, Raveh A,
Dinamarca MC, Hanuschkin A, Bildl W, Klingauf J, Gassmann M, Schulte U, Bettler B, Fakler
B (2016) Modular composition and dynamics of native GABAB receptors identified by high-
resolution proteomics. Nat Neurosci 19(2):233–242
Simonian SX, Skynner MJ, Sieghart W, Essrich C, Luscher B, Herbison AE (2000) Role of the
GABA(A) receptor gamma2 subunit in the development of gonadotropin-releasing hormone
neurons in vivo. Eur J Neurosci 12(10):3488–3496
Stratton MS, Searcy BT, Tobet SA (2011) GABA regulates corticotropin releasing hormone levels
in the paraventricular nucleus of the hypothalamus in newborn mice. Physiol Behav
104(2):327–333
Stratton MS, Staros M, Budefeld T, Searcy BT, Nash C, Eitel C, Carbone D, Handa RJ, Majdic G,
Tobet SA (2014) Embryonic GABA(B) receptor blockade alters cell migration, adult hypotha-
lamic structure, and anxiety- and depression-like behaviors sex specifically in mice. PLoS One
9(8):e106015
Suter KJ, Wuarin JP, Smith BN, Dudek FE, Moenter SM (2000) Whole-cell recordings from
preoptic/hypothalamic slices reveal burst firing in gonadotropin-releasing hormone neurons
identified with green fluorescent protein in transgenic mice. Endocrinology 141(10):3731–3736
Temple JL, Wray S (2005) Developmental changes in GABA receptor subunit composition within
the gonadotrophin-releasing hormone-1 neuronal system. J Neuroendocrinol 17(9):591–599
Tiao J, Bettler B (2007) Characteristics of GABAB receptor mutant mice. In: The GABA receptors,
3rd edn. Springer, Berlin
Tobet SA, Paredes RG, Chickering TW, Baum MJ (1995) Telencephalic and diencephalic origin of
radial glial processes in the developing preoptic area/anterior hypothalamus. J Neurobiol
26(1):75–86
Tobet SA, Chickering TW, King JC, Stopa EG, Kim K, Kuo-Leblank V, Schwarting GA (1996a)
Expression of gamma-aminobutyric acid and gonadotropin-releasing hormone during neuronal
migration through the olfactory system. Endocrinology 137(12):5415–5420
Tobet SA, Hanna IK, Schwarting GA (1996b) Migration of neurons containing gonadotropin
releasing hormone (GnRH) in slices from embryonic nasal compartment and forebrain. Brain
Res Dev Brain Res 97(2):287–292
Tobet SA, Henderson RG, Whiting PJ, Sieghart W (1999) Special relationship of gamma-
aminobutyric acid to the ventromedial nucleus of the hypothalamus during embryonic develop-
ment. J Comp Neurol 405(1):88–98
Tobet SA, Bless EP, Schwarting GA (2001) Developmental aspect of the gonadotropin-releasing
hormone system. Mol Cell Endocrinol 185(1–2):173–184
Tobet SA, Walker HJ, Seney ML, Yu KW (2003) Viewing cell movements in the developing
neuroendocrine brain. Integr Comp Biol 43(6):794–801
Tobet SA, Handa RJ, Goldstein JM (2013) Sex-dependent pathophysiology as predictors of
comorbidity of major depressive disorder and cardiovascular disease. Pflugers Arch
465(5):585–594
Vastagh C, Schwirtlich M, Kwakowsky A, Erdélyi F, Margolis FL, Yanagawa Y, Katarova Z,
Szabó G (2015) The spatiotemporal segregation of GAD forms defines distinct GABA signaling
functions in the developing mouse olfactory system and provides novel insights into the origin
and migration of GnRH neurons. Dev Neurobiol 75(3):249–270
Wierman ME, Kiseljak-Vassiliades K, Tobet S (2011) Gonadotropin-releasing hormone (GnRH)
neuron migration: initiation, maintenance and cessation as critical steps to ensure normal
reproductive function. Front Neuroendocrinol 32(1):43–52
7 Hypothalamic Development: Role of GABA 205
Abstract
During the last few years, new genes required for the advent of puberty have been
identified thanks to the use of new deep sequencing associated technologies.
Evidence has emerged suggesting that these genes, organized into functional
networks, are regulated by a mechanism of epigenetic control. While the
Polycomb group (PcG) of transcriptional repressors is in charge of the repressive
arm of this mechanism, the Trithorax Group (TrxG) of epigenetic activators
counteract the PcG-mediated silencing during the infantile to juvenile stage
where increased Kiss1 expression is needed for activation of gonadotropin-
releasing hormone secretion from the hypothalamus. Here we will discuss the
recent advances and tools used for large-scale genomic and epigenomic studies of
the developing hypothalamus.
Keywords
Epigenetic regulators · Female puberty gene networks · GnRH neurons · KNDy
neurons · Reproductive development · Transcriptional activators · Transcriptional
repressors
8.1 Introduction
GnRH neurons are responsible for the increased GnRH secretion. These changes
involve a reduction in inhibitory trans-synaptic inputs combined with increased
trans-synaptic and glial excitatory inputs to the GnRH neuronal network (Ojeda
and Skinner 2006).
The time of puberty has a strong genetic component, but very recently epigenetic
mechanisms have been implicated as part of the regulatory control of the develop-
mental process by which GnRH release is kept in check during infancy, and also
during the increase in GnRH secretion that brings about the pubertal process.
Agreeing with this emergent concept, a central target of epigenetic control is the
transcriptional machinery of neurons in charge of stimulating GnRH release. In this
chapter, we review key findings as well as research methodologies used to support
the existence of an epigenetically controlled gene regulatory network residing at the
core of the neuroendocrine process that controls pubertal timing.
With the advent of new high-throughput genomic technologies, coupled with a novel
and developing systems biology tools, great advances have been made in determin-
ing new pathways involved in the neuroendocrine control of puberty. This novel
approach is based on the concept that a diversity of genes that affect the time of
puberty are functionally organized into networks capable of generating a coordinated
output of biological signals. About 10 years ago, using expression data from female
rats and monkeys, we described the first gene network operating in the peripubertal
hypothalamus. This gene network is organized into a structure containing “central”
nodes that reside at the core of the network, and a number of peripherally located
subordinate genes that are transcriptionally controlled by the central nodes. The
central nodes of this network were earlier identified as being implicated in tumor
suppression/tumor formation (Roth et al. 2007), now referred to as Tumor-Related
Genes (TRGs). The TRG network has five central nodes or “hubs”: CDP/CUTL1/
CUX1, MAF, p53, YY1, and USF2. These hubs are strongly connected to each other
and also to additional upper-echelon genes (OCT2, TTF1, and EAP1) previously
found to be involved in the transcriptional regulation of the pubertal process (Heger
et al. 2007; Mastronardi et al. 2006; Ojeda et al. 1999), and TRGs themselves
(Lomniczi et al. 2013b).
We (Heger et al. 2007; Lomniczi et al. 2012; Mueller et al. 2011, 2012) and others
(Cukier et al. 2013) have experimentally verified the role of EAP1 as a central
transcriptional regulator of subordinate genes like KISS1 and GnRH. Interestingly,
KISS1 was formerly known as a Metastasis Suppressor or Metastin (Steeg et al.
2003), while EAP1 was described as part of a transcriptional repressive complex that
regulates apoptosis in breast cancer (Yeung et al. 2011). Most recently, we found a
210 C. F. Aylwin and A. Lomniczi
The first step into analyzing RNA expression or chromatin modifications throughout
critical developmental periods is to characterize the phenotype of interest, link it to a
specific day of pre- or postnatal development and to a tissue or brain region. In the rat
model, the Hypothalamus–Pituitary–Gonadal (HPG) axis remains quiescent until
postnatal day 28 (PND28) when the first endocrine manifestations of increased
GnRH pulsatility are detected as LH pulse frequency increases during the evening
hours (Urbanski and Ojeda 1985). After a few days of increased diurnal LH
pulsatility, and around PND32–34, the first preovulatory surge of gonadotropins
occurs, followed by the first ovulation. In general terms, pubertal development can
be divided into three stages: juvenile period (Juv ¼ PND21), late juvenile
(LJ ¼ PND28), and late proestrus (LP ¼ PND32–34) (Fig. 8.1).
In the rhesus macaque, we define four different stages of reproductive develop-
ment: neonatal, infantile, juvenile, and pubertal. As in humans, during the first
months to a year of life, high levels of LH pulsatility are detected in a period called
“mini-puberty” This is followed by the infantile period in which there is a marked
quiescence of the gonadotropin axis. During the early pubertal period, there is a
Fig. 8.1 Plasma LH concentrations during postnatal development in humans, rhesus macaques,
and rats. In humans and monkeys, during the early postnatal period, high LH pulsatility is detected
in a period called “mini-puberty.” This is followed by the infantile period where a “repressed”
central drive generates low LH levels. During the juvenile stage, 8–9 years of age in humans and
approximately 2 years of age for the rhesus macaques, there is a reawakening of the GnRH network
evidenced by increased LH pulsatility during the night hours. In the rodent, this same phenomenon
is first detected in the evening of postnatal days 28–29. As puberty approaches in females, ovarian
estrogen secretion increases, reaching a sustained magnitude capable of activating the positive
feedback mechanism that results in the preovulatory surge of gonadotropins. Days (d), months (m),
and years (yrs)
8 Epigenetic and Transcriptional Regulation of the Reproductive Hypothalamus 213
Fig. 8.2 Schematic representation of the hypothalamic region of the rat. (a) Ventral view of the rat
brain. Major regions (olfactory bulb, neocortex, optic chiasma, hypothalamus) are labeled. The
hypothalamic region, shaded in orange, is used for micro-punching or coronal slicing. (b) Coronal
view of the hypothalamic region depicting the arcuate nucleus (Arc), median eminence (ME),
ventromedial hypothalamus (VMH), and dorsomedial hypothalamus (DMD)
214
Fig. 8.3 RNA-Seq library preparation process from rat MBH tissue. (a) Agilent Bioanalyzer electrophoretic trace of a high-quality total RNA sample. Shaded
green area shows the 18S and 28S rRNA bands. (b) Electrophoretic trace after rRNA depletion using poly-A magnetic beads. The green shaded area highlights
the absence of 18S and 28S signals depicted by dotted lines. The blue line corresponds to the purified mRNA 1000–5000 nt. (c) Electrophoretic trace after
chemical fragmentation of mRNA. Fragmented mRNA can be observed by a shift of the enriched area toward the lower molecular weight, 100–800 nt (orange
line) as opposed to intact mRNA (dotted blue line). (d) Electrophoretic trace of a finalized sequencing library after amplification using primers P5 and P7
followed by magnetic beads size selection to discard both adapter and primer dimers. An enrichment in 400–600 bp long amplicons is detected
C. F. Aylwin and A. Lomniczi
8 Epigenetic and Transcriptional Regulation of the Reproductive Hypothalamus 215
of ribosomal RNA using either active ribosome binding technology like Ribo-Zero
rRNA removal kit (Illumina cat# MRZH116) or by further isolating the poly-A
mRNA from the total RNA fraction by means of poly-(A) paramagnetic beads (Bioo
Scientific Cat# NOVA-51979). These magnetic beads are coupled to oligo
(dT) probes capable of isolating poly-A mRNA starting from 10 ng to 10 μg of
total RNA (See Box 8.2). Poly-A beads are more price effective than any of the
ribosomal depletion methods, with the caveat of generating libraries that only
represent poly-A containing mRNA. The first step is to test the best condition for
mRNA isolation. One efficient way to do this is to use equal volumes of poly-(A)
beads with increasing amounts of total RNA. Efficient removal of ribosomal RNA
can be identified by the absence of 18S and 28S fractions commonly seen in total
RNA samples in Bioanalyzer RNA nano- or pico chip runs (Agilent cat# 5067-1511
and 5067-1514) (Fig. 8.3b).
RNA transcripts can be isolated from total RNA by two different methods:
(a) Ribosomal depletion, a method where all ribosomal RNAs are captured and
discarded from the sample, allows for isolation of both mRNA and lncRNA
and (b) Poly-A-enrichment, specifically captures and purifies poly-A-
containing mRNAs using poly-T-containing magnetic beads. After isolation,
samples are subjected to fragmentation and cDNA synthesis. Adapter ligation
allows for the attachment of Illumina compatible adapter sequences and
barcodes for downstream sequencing and sample identification. Before
sequencing, library quantification by qPCR is necessary for accuracy.
216 C. F. Aylwin and A. Lomniczi
Several Illumina compatible kits for RNA-seq library preparation are available on
the market. In general, the basic steps to prepare a next-generation sequencing
(NGS) library include (a) fragmentation of the RNA sample; (b) converting RNA
into double-stranded DNA; (c) attaching oligonucleotide adapters to each end of the
target fragments; and (d) quantitating the library by qPCR. The first step of this
process is RNA fragmentation and can be achieved by (a) physical fragmentation
like acoustic or hydrodynamic shearing; (b) enzymatic shearing like RNase III
treatment; and (c) chemical fragmentation by incubating the mRNA sample at
95 C with divalent cations. The optimal size fragmentation for Illumina compatible
libraries ranges from 100 to 400 nt, this can also be detected by Bioanalyzer RNA
nano- or pico chip (Fig. 8.3c). After fragmentation, first strand synthesis is
performed with random primers, if it is desired to retain the strand orientation of
the transcripts, dUTP is added into the second strand synthesis, which prevents
further amplification with the PCR polymerase that is unable to recognize uracil in
the template strand (NEXTflex Rapid Directional RNA-seq kit, Bioo Scientific cat#
NOVA-5138-07). Bead cleanup using Agencourt AMPure XP Magnetic Beads is
performed in order to remove free oligonucleotides and dNTPs from the reaction
mix. This is followed by 30 end adenylation and adapter ligation. The Illumina
compatible adapters carry the barcodes used to identify the sample after sequencing
as well as the primer recognition sites used for library amplification as well as
sequencing on the Illumina flow cell. After adapter ligation, limited cycle PCR
amplification of the library is performed in order to reach the minimum amount of
DNA needed for sequencing. Size selection is performed on the library with AMPure
XP Magnetic beads, this step is needed to remove the primers from the mix that
could interfere with the sequencing reaction (Fig. 8.3d). The library is quantitated by
qPCR using the Kapa Library Quantitation Kit (Roche cat# 07960140001). Six to
eight samples with different barcodes are pooled into a single tube at a concentration
of 3 nM and send for sequencing in an Illumina HiSeq 3000 or 4000 sequencer that
delivers around 300 million reads per lane, providing about 50 million reads per
sample when using pools of six samples per sequencing lane.
The Illumina platform delivers raw reads in fasqc formatted files. The first step is to
run quality controls on the sequencing procedure, followed by alignment of the
sequencing reads to a reference genome and quantitation of reads per gene. All
the computational tools needed to analyze NGS datasets are freely available at the
Bioconductor webpage (http://www.bioconductor.org/), and are coded on R lan-
guage (https://www.r-project.org/). Another platform widely used for high-
throughput data analysis is Galaxy (https://usegalaxy.org/), which is especially
helpful for users with less experience in programming on R. Finally, DNASTAR
(https://www.dnastar.com/) is a commercial package that fulfills all the steps from
quality control, data alignment, transcript quantitation as well as data visualization
and gene ontology analysis.
8 Epigenetic and Transcriptional Regulation of the Reproductive Hypothalamus 217
Fig. 8.4 CHIP-Seq library preparation process from rat MBH tissue. (a) In order to determine the sonication time to generate optimum chromatin
219
fragmentation, the tissue is treated with increasing number of pulses followed by de-crosslinking, DNA purification, and gel electrophoresis. Increasing the
220
Fig. 8.4 (continued) number of pulses reduces the average size of the fragments. Large fragments (>500 bp) reduce the accuracy of the CHIP procedure. The
optimum fragment size is between 150 and 500 bp. (b) Linearity of chromatin capture is determined by using a constant amount of Antibody (2–5 ug) and
increasing amounts of chromatin. After the CHIP assay is completed, cycle threshold (ct) values amplification of a specific loci by qPCR must decrease linearly
according to the amount of chromatin used (blue circles corresponding to CHIPed H3K27me3). The green shaded area shows the linear range of chromatin/
antibody interaction. The red and yellow areas depict saturated chromatin/antibody conditions where the qPCR is not capable to detect changes in DNA capture.
The orange circles correspond to an anti beta-Galactosidase antibody (βGAL) used as negative control showing no efficient chromatin/antibody interactions. (c)
Agilent Bioanalyzer electrophoretic trace of an amplified chromatin library generates a signal corresponding to the barcoded library (green shaded area). The low
molecular weight peak corresponds to adapter dimers (red shaded area), while the high molecular weight peaks are considered artifacts due to overamplification
of the library (yellow shaded area). (d) Electrophoretic trace of an amplified chromatin library after double size selection to discard both low- and high-molecular
peaks. First, and to cleanup large molecular weight fragments, a high magnetic bead:library ratio (0.8–1) is used to bind large DNA fragments while the library
and the small fragments remain unbound. The soluble unbound DNA is transferred into a new tube and a low magnetic bead:library ratio (0.2) is used, the
desired library is captured by the beads, while the small fragments remain unbound. After several washes, the bound library fragments are eluted
C. F. Aylwin and A. Lomniczi
8 Epigenetic and Transcriptional Regulation of the Reproductive Hypothalamus 221
Magnetic beads before PCR amplification of the library, to remove free unligated
adapters and high molecular weight DNA that has been selected during immunopre-
cipitation (Fig. 8.4c, d). To correct for technical variability during CHIP/Library
preparation, it is suggested to add a spike-in control to each CHIP sample (Egan et al.
2016). We usually add 1% of chromatin from Drosophila and antibodies to immu-
noprecipitate histone H2Av, only present in this species. Since this step will pull
down the same amount of Drosophila histone 2Av between samples, it allows for
normalization of the CHIP-seq values from the mammalian genome to the total
number of reads obtained from the Drosophila genome.
Platforms used for CHIP-seq analysis are the same as for RNA-seq. It is estimated
that a reading depth of about 20 million reads per sample is adequate to detect
transcription factor binding in the mammalian genome. Proteins with more binding
sites as well as histone marks require up to 60 million reads. After assessment of the
reads quality metrics, a mapping tool like Bowtie (Langmead et al. 2009) or MAQ
(Li and Durbin 2009) is used to align the short reads to the genome. The analysis is
followed by peak calling algorithms like MACS (Zhang et al. 2008) that predict the
regions of the genome where the immunoprecipitated protein is bound. Peak callers
will provide a p-value and FDR as enrichment metrics for each peak. For enzyme or
transcription factor binding, peaks are readily detected, but in the case of histone
PTMs where marks are more broadly positioned, it is more accurate to define
“chromatin domains.” Quantitative analysis is based on the total read count per
region of interest. DBCHIP (Liang and Keles 2012) and MAnorm (Shao et al. 2012)
compute read density over peak regions and give a precise statistical assessment of
differentially affected regions across biological conditions. Peaks or domains
identified by CHIP-seq are coded as BED or GFF files that can be uploaded to a
genome browser for further identification of promoter and enhancer regions as well
as gene transcription start sites.
increases, lowering their expression, inducing the loss of PcG components from the
Kiss1 promoter. This is accompanied by increased levels of the epigenetic marks
associated with gene activation like H3K9ac, H3K14ac, and H3K4me3, ultimately
inducing increased Kiss1 mRNA expression in the ARC (Lomniczi et al. 2013a). In a
second paper, and using rhesus macaques as animal model, we provided evidence
that members of the zinc finger (ZNF) family of transcriptional repressors play a role
in keeping the GnRH pulse generator in check during prepubertal development
(Lomniczi et al. 2015). Using MBH of both ovary-intact females and agonadal
male rhesus monkeys during the juvenile–pubertal transition we demonstrated that
while KISS1 and TAC3 mRNA expression increases, GATAD1 expression decreases.
GATAD1 recruits the histone eraser KDM1A, inducing a decrease in the activating
H3K4me3/2 marks at the Kiss1 and TAC3 genes 50 -regulatory regions (Lomniczi
et al. 2015; Maisonpierre et al. 1991; Shi et al. 2004). Stereotaxic delivery of
lentiviral particles overexpressing GATAD1 in the ARC of immature rats, delays
puberty, and disrupts estrous cyclicity (Lomniczi et al. 2015). The decrease in
GATAD1 and KDM1A association to the KISS1 and TAC3 promoters and the
concomitant increase in H3K4me2 abundance observed in the monkey MBH at
the time of puberty, strongly support the notion that during juvenile development
there is a mechanism of epigenetic repression, that is lifted during the reawakening
of the GnRH pulse generator (Lomniczi et al. 2015).
The finding that the chromatin landscape at the Kiss1 promoter switches from a
primarily repressed state, brought about by the PcG and GATAD1/KDM1A com-
plex, to an activating state characterized by trimethylation and acetylation of H3 at
lysines 4 and 9, respectively, during the peri-pubertal period indicates the action of
an activating transcriptional complex occurring simultaneously to the loss of inhibi-
tion. This idea was studied by RNA-seq analysis of the rat MBH during peri-pubertal
development. In this study, we found that members of the Trithorax Group (TrxG) of
transcriptional activators, well known for antagonizing PcG activity (Shilatifard
2012; Simon and Kingston 2009), increase their expression in the MBH as puberty
progresses (Toro et al. 2018). MLL1 and MLL3, two members of the TrxG complex,
act on the Kiss1 promoter and enhancer regions, respectively, to provide trans-
activational activity at the time when the inhibitory influence of the PcG complex is
declining (Toro et al. 2018). To confirm TrxG function, siRNA-mediated knock-
down of Mll1 in the ARC, inhibited Kiss1 expression and delayed puberty (Toro
et al. 2018).
Since the TrxG member MLL3 has a preference for binding at enhancer sites, and
active enhancers display high levels of H3K4me1 and H3K27ac (Kim and
Shiekhattar 2015), we performed CHIP-seq analysis of these two histone
modifications in the MBH of prepubertal rats to pinpoint potential enhancers of
the ARC Kiss1 gene. In a 12-kb segment of the Kiss1 gene 50 upstream region, we
detected two areas showing high levels of H3K4me1 and H3K27ac. CHIP-qPCR
analysis of Site 1 and Site 2 in the rat genome revealed that only Site 1 (about
3000 bp from TSS) displayed increased association of p300/CBP and H3K27ac
during prepubertal development, a hallmark of active enhancers (Kim et al. 2010;
Kim and Shiekhattar 2015). In agreement with this concept, MLL3 recruitment to
8 Epigenetic and Transcriptional Regulation of the Reproductive Hypothalamus 223
Site 1 (but not Site 2) increased during pubertal development, while the
EED-mediated H3K27m3 mark weans down (Toro et al. 2018). In order to deter-
mine the function of this new Kiss1 enhancer region, we used a CRISPR-CAS9-
KRAB epigenetic remodeling approach that inhibits the action of Mll3 by site-
specific deposition of the repressive H3K9m3 histone mark. Animals stereotaxically
injected with an AAV9 virus carrying this construct into the ARC experienced a
downregulation of Kiss1 expression, delayed puberty and disrupted estrous cyclicity
(Toro et al. 2018). In conclusion, the TrxG of epigenetic activators is capable of
controlling Kiss1 expression in the ARC, by implementing histone PTM changes at
promoter and enhancer regions around the time of puberty.
the critical mass needed for library preparation and sequencing. While single-cell
mRNA sequencing technologies have been developed (Eberwine et al. 2014),
allowing for picogram amounts of starting material to be used, heavy amplification
is needed for sample and library preparation, resulting in uneven coverage, high
noise, and inaccurate quantitation. Needless to say, CHIP-seq technologies have
improved tremendously, but efficiency and sequence quality get compromised when
starting material lies below 104 cells (Gilfillan et al. 2012).
The second limitation is related to the massive amount of information generated
and the wide range of platforms and algorithms that can be used to analyze and
visualize the data. Although the ENCODE Consortium (https://www.encodeproject.
org/) has set the standard for the analysis of most of these technologies, researchers
have to rely on external bioinformatics cores, recruit informaticians in their
laboratories or even learn coding and statistics for high-throughput data analysis.
This can be especially challenging for small laboratories that do not have easy access
to core facilities and other informatics resources.
With the widespread use of NGS technologies, scientists struggle to find ways to
identify the relevant genes or genome signatures from long gene lists. For instance,
genes are scored by their differential expression or chromatin signatures between
different biological states. Using an arbitrary expression threshold (we usually use
1.5-fold expression change), gene lists are screened for the top most variable genes.
Searching for enrichment of predefined functionally related genes can be easily done
with stand-alone software like Ingenuity Pathway Analysis (IPA) from Qiagen
(https://www.qiagenbioinformatics.com/products/ingenuity-pathway-analysis/),
web-based program like DAVID Bioinformatic Database (https://david.ncifcrf.gov/
home.jsp) (Dennis et al. 2003) and GeneMANIA (https://genemania.org/) (Mostafavi
et al. 2008) or open-source options like several plugins for Cytoscape (http://www.
cytoscape.org/) (Cline et al. 2007).
Fig. 8.5 Gene ontology (GO) analysis of the genes with SNPs associated with age at menarche.
DAVID (https://david.ncifcrf.gov/home.jsp) GO categories: Biological Process and Molecular
Function of the 339 genes with SNPs associated with the age at menarche show a high enrichment
in, e.g., gene transcription, chemical synapses, steroid signaling, and others
Fig. 8.6 Gene network representation of the genes with SNPs associated with age at menarche and
their GeneMANIA connections with the epigenetic machinery. The round nodes of the network are
the 339 genes associated with age at menarche, red diamonds are genes of the Polycomb Group
(PcG) of transcriptional repressors, orange diamonds represent Gatad1 and Kdm1a and green
squares are the Trithorax Group (TrxG) of transcriptional activators. The epigenetic machinery
(depicted on the first layer of the network) is connected to 229 out of 339 of the genes associated
with age at menarche (second layer of nodes). The third layer depicts the 110 genes that are not
directly connected to the epigenetic machinery. Yellow nodes are genes involved in neurotransmis-
sion, blue nodes are transcriptional repressors and cyan nodes are transcriptional activators. Gray
lines, or edges, are connections established by the GeneMANIA database representing: Genetic
interactions, Pathways, Physical interactions, Predicted, and Shared protein domains
the SNP’s described by Day et al. (2017) as associated with the age at menarche, are
in some way altering the epigenetic milieu of their downstream target genes.
Fig. 8.7 Effect of GATAD1 overexpression on gene co-expression networks in the hypothalamus
of rhesus macaques. (a) Inferred co-expression network between the top 20 genes losing edges
(green) and the top 20 genes gaining edges/connections (red) during prepubertal development,
based on the sum of changes in connectivity between PND14–PND21 and PND21–PND28.
Network at PND 28 with the top 20 genes losing edges (yellow) and gaining edges (blue) under
8 Epigenetic and Transcriptional Regulation of the Reproductive Hypothalamus 229
demonstration that during the infantile period, when Kiss1 expression is low, the
GATAD1/LSD1 complex is associated to the Kiss1 promoter keeping low levels of
H3K4me3. As puberty approaches and Gatad1 expression decreases, the GATAD1/
LSD1 complex is evicted from the promoter, H3K4me2/3 rises, and Kiss1 expres-
sion increases (Lomniczi et al. 2015).
One of the biggest limitations is related to the animal species of choice. Most of the
interactomics databases is limited to human and/or mouse datasets. During the last
few years, efforts have been made to widen this repertoire to include rats, Drosoph-
ila, C. elegans, and others. With that in mind, researchers many times opt to use
human or mouse databases when investigating rat genomic interactions due to the
higher characterization of gene function and interactions in these species. Another
issue that comes out frequently is that not all genes are yet classified on the Gene
Ontology database, producing bias toward a subclass of previously well-studied
genes present in the dataset.
Network analysis and visualization are powerful tools for whole-genome studies,
but it is crucial for investigators to take all the precautions for the correct statistical
analysis and subsequent validation of the identified pathways or gene interactions.
8.7 Perspectives
Fig. 8.7 (continued) GATAD1 overexpression as compared with the network inferred at PND28.
In this network, genes that are among the top 20 genes gaining edges during normal prepubertal
development are shown as triangles, while genes that are among the top 20 losers of connectivity
during development are shown as squares. For all networks, positive correlation edges are
represented by solid black lines, negative correlation edges are shown as dashed lines. (b) Changes
in the connectivity of genes belonging to selected transcription-related functional categories that
lose (Gene Ontology category GO:0045892: negative regulation of transcription, DNA templated)
or gain (Protein Information Resource, http://pir.georgetown.edu/; Keyword “transcription regula-
tion”) connectivity (green and red bars, respectively) from PND14 to PND28. For each gene, the
alteration in connectivity resulting from GATAD1 overexpression as compared with PND28 is
shown as blue bars. Modified from Lomniczi et al. (2015)
230 C. F. Aylwin and A. Lomniczi
Acknowledgments We thank every past member of our laboratory who has contributed to the
understanding of the genomic events involved in the maturation of the reproductive hypothalamus.
The authors are supported by the National Institute of Health (1R01HD084542 and
P51OD011092).
Key References
Boyar et al. (1972). This study was the first demonstrating that LH pulsatility
increases before sleep in pubertal patients.
Lomniczi et al. (2013). This study describes how the Polycomb family of epigenetic
repressors control the epigenetic landscape at the Kiss1 promoter region, keeping
ARC Kiss1 transcription low during the infantile period.
Lomniczi et al. (2015). This study describes how the GATAD1/LSD1 controls the
epigenetic landscape at the Kiss1 promoter region in rhesus macaques, keeping
ARC Kiss1 transcription low during the infantile period.
Roth et al. (2007). This study was the first proposing a tumor-related gene network
that controls mammalian pubertal development.
Toro et al. (2018). This study describes how the Trithorax family of epigenetic
activators control the epigenetic landscape at the Kiss1 promoter/enhancer
regions during the pubertal activation on Kiss1 transcription in the ARC.
References
Amar D, Safer H, Shamir R (2013) Dissection of regulatory networks that are altered in disease via
differential co-expression. PLoS Comput Biol 9:E1002955
Anders S, Huber W (2010) Differential expression analysis for sequence count data. Genome Biol
11:R106
Barutcu AR, Fritz AJ, Zaidi SK, Van Wijnen AJ, Lian JB, Stein JL, Nickerson JA, Imbalzano AN,
Stein GS (2016) C-Ing the genome: a compendium of chromosome conformation capture
methods to study higher-order chromatin organization. J Cell Physiol 231:31–35
Beale KE, Kinsey-Jones JS, Gardiner JV, Harrison EK, Thompson EL, Hu MH, Sleeth ML, Sam
AH, Greenwood HC, McGavigan AK, Dhillo WS, Mora JM, Li XF, Franks S, Bloom SR,
O'Byrne KT, Murphy KG (2014) The physiological role of arcuate kisspeptin neurons in the
control of reproductive function in female rats. Endocrinology 155(3):1091–1098
Bindea G, Mlecnik B, Hackl H, Charoentong P, Tosolini M, Kirilovsky A, Fridman WH, Pages F,
Trajanoski Z, Galon J (2009) Cluego: a cytoscape plug-in to decipher functionally grouped gene
ontology and pathway annotation networks. Bioinformatics 25:1091–1093
Bolger AM, Lohse M, Usadel B (2014) Trimmomatic: a flexible trimmer for illumina sequence
data. Bioinformatics 30:2114–2120
Boyar R, Finkelstein J, Roffwarg H, Kapen S, Weitzman E, Hellman L (1972) Synchronization of
augmented luteinizing hormone secretion with sleep during puberty. N Engl J Med 287:582–
8 Epigenetic and Transcriptional Regulation of the Reproductive Hypothalamus 231
586. [This study was the first demonstrating that LH pulsatility increases before sleep in pubertal
patients]
Buenrostro JD, Wu B, Chang HY, Greenleaf WJ (2015) Atac-seq: a method for assaying chromatin
accessibility genome-wide. Curr Protoc Mol Biol 109:21.29.1–21.29.9
Chang TC, Zeitels LR, Hwang HW, Chivukula RR, Wentzel EA, Dews M, Jung J, Gao P, Dang
CV, Beer MA, Thomas-Tikhonenko A, Mendell JT (2009) Lin-28B transactivation is necessary
for Myc-mediated let-7 repression and proliferation. Proc Natl Acad Sci U S A 106(9):3384–
3389
Cline MS, Smoot M, Cerami E, Kuchinsky A, Landys N, Workman C, Christmas R, Avila-
Campilo I, Creech M, Gross B, Hanspers K, Isserlin R, Kelley R, Killcoyne S, Lotia S,
Maere S, Morris J, Ono K, Pavlovic V, Pico AR, Vailaya A, Wang PL, Adler A, Conklin
BR, Hood L, Kuiper M, Sander C, Schmulevich I, Schwikowski B, Warner GJ, Ideker T, Bader
GD (2007) Integration of biological networks and gene expression data using cytoscape. Nat
Protoc 2:2366–2382
Crawford GE, Holt IE, Whittle J, Webb BD, Tai D, Davis S, Margulies EH, Chen Y, Bernat JA,
Ginsburg D, Zhou D, Luo S, Vasicek TJ, Daly MJ, Wolfsberg TG, Collins FS (2006) Genome-
wide mapping of dnase hypersensitive sites using massively parallel signature sequencing
(Mpss). Genome Res 16:123–131
Cukier P, Wright H, Rulfs T, Silveira LF, Teles MG, Mendonca BB, Arnhold IJ, Heger S, Latronico
AC, Ojeda SR, Brito VN (2013) Molecular and gene network analysis of thyroid transcription
factor 1 (TTF1) and enhanced at puberty (EAP1) genes in patients with GnRH-dependent
pubertal disorders. Horm Res Paediatr 80(4):257–266
D’haeseleer P (2005) How does gene expression clustering work? Nat Biotechnol 23:1499–1501
Day FR, Thompson DJ, Helgason H, Chasman DI, Finucane H, Sulem P, Ruth KS, Whalen S,
Sarkar AK, Albrecht E, Altmaier E, Amini M, Barbieri CM, Boutin T, Campbell A,
Demerath E, Giri A, He C, Hottenga JJ, Karlsson R, Kolcic I, Loh PR, Lunetta KL,
Mangino M, Marco B, Mcmahon G, Medland SE, Nolte IM, Noordam R, Nutile T,
Paternoster L, Perjakova N, Porcu E, Rose LM, Schraut KE, Segre AV, Smith AV, Stolk L,
Teumer A, Andrulis IL, Bandinelli S, Beckmann MW, Benitez J, Bergmann S, Bochud M,
Boerwinkle E, Bojesen SE, Bolla MK, Brand JS, Brauch H, Brenner H, Broer L, Bruning T,
Buring JE, Campbell H, Catamo E, Chanock S, Chenevix-Trench G, Corre T, Couch FJ,
Cousminer DL, Cox A, Crisponi L, Czene K, Davey Smith G, De Geus E, De Mutsert R, De
Vivo I, Dennis J, Devilee P, Dos-Santos-Silva I, Dunning AM, Eriksson JG, Fasching PA,
Fernandez-Rhodes L, Ferrucci L, Flesch-Janys D, Franke L, Gabrielson M, Gandin I, Giles GG,
Grallert H, Gudbjartsson DF, Guenel P, Hall P, Hallberg E, Hamann U, Harris TB, Hartman CA,
Heiss G, Hooning MJ, Hopper JL, Hu F, Hunter DJ, Ikram MA, Im HK, Jarvelin MR, Joshi PK,
Karasik D, Kellis M et al (2017) Genomic analyses identify hundreds of variants associated with
age at menarche and support a role for puberty timing in cancer risk. Nat Genet 49:834–841
de Roux N, Genin E, Carel J-C, Matsuda F, Chaussain J-L, Milgrom E (2003) Hypogonadotropic
hypogonadism due to loss of function of the KiSS1-derived peptide receptor GPR54. Proc Natl
Acad Sci U S A 100:10972–10976
Dennis G Jr, Sherman BT, Hosack DA, Yang J, Gao W, Lane HC, Lempicki RA (2003) David:
database for annotation, visualization, and integrated discovery. Genome Biol 4:P3
Dobin A, Davis CA, Schlesinger F, Drenkow J, Zaleski C, Jha S, Batut P, Chaisson M, Gingeras TR
(2013) Star: ultrafast universal RNA-seq aligner. Bioinformatics 29:15–21
Eberwine J, Sul JY, Bartfai T, Kim J (2014) The promise of single-cell sequencing. Nat Methods
11:25–27
Egan B, Yuan CC, Craske ML, Labhart P, Guler GD, Arnott D, Maile TM, Busby J, Henry C, Kelly
TK, Tindell CA, Jhunjhunwala S, Zhao F, Hatton C, Bryant BM, Classon M, Trojer P (2016) An
alternative approach To ChIP-seq normalization enables detection of genome-wide changes in
histone h3 lysine 27 trimethylation upon EZH2 inhibition. PLoS One 11:E0166438
Friedman N, Linial M, Nachman I, Pe'er D (2000) Using Bayesian networks to analyze expression
data. J Comput Biol 7:601–620
Garcia-Martinez J, Aranda A, Perez-Ortin JE (2004) Genomic run-on evaluates transcription rates
for all yeast genes and identifies gene regulatory mechanisms. Mol Cell 15:303–313
232 C. F. Aylwin and A. Lomniczi
Gilfillan GD, Hughes T, Sheng Y, Hjorthaug HS, Straub T, Gervin K, Harris JR, Undlien DE, Lyle
R (2012) Limitations and possibilities of low cell number ChIP-seq. BMC Genomics 13:645
Giresi PG, Kim J, Mcdaniell RM, Iyer VR, Lieb JD (2007) Faire (formaldehyde-assisted isolation
of regulatory elements) isolates active regulatory elements from human chromatin. Genome Res
17:877–885
Glass K, Huttenhower C, Quackenbush J, Yuan GC (2013) Passing messages between biological
networks to refine predicted interactions. PLoS One 8:E64832
Hagan JP, Piskounova E, Gregory RI (2009) Lin 28 recruits the TUTase Zcchc11 to inhibit let-7
maturation in mouse embryonic stem cells. Nat Struct Mol Biol 16(10):1021–1025
Heger S, Mastronardi C, Dissen GA, Lomniczi A, Cabrera R, Roth CL, Jung H, Galimi F,
Sippell W, Ojeda SR (2007) Enhanced at puberty 1 (EAP1) is a new transcriptional regulator
of the female neuroendocrine reproductive axis. J Clin Invest 117(8):2145–2154
Heo I, Joo C, Kim YK, Ha M, Yoon MJ, Cho J, Yeom KH, Han J, Kim VN (2009) TUT4 in concert
with Lin28 suppresses microRNA biogenesis through pre-microRNA uridylation. Cell 138
(4):696–708
Kim TK, Shiekhattar R (2015) Architectural and functional commonalities between enhancers and
promoters. Cell 162(5):948–959
Kim TK, Hemberg M, Gray JM, Costa AM, Bear DM, Wu J, Harmin DA, Laptewicz M, Barbara-
Haley K, Kuersten S, Markenscoff-Papadimitriou E, Kuhl D, Bito H, Worley PF, Kreiman G,
Greenberg ME (2010) Widespread transcription at neuronal activity-regulated enhancers.
Nature 465(7295):182–187
Kwak H, Fuda NJ, Core LJ, Lis JT (2013) Precise maps of RNA polymerase reveal how promoters
direct initiation and pausing. Science 339:950–953
Langmead B, Trapnell C, Pop M, Salzberg SL (2009) Ultrafast and memory-efficient alignment of
short DNA sequences to the human genome. Genome Biol 10:R25
Law CW, Chen Y, Shi W, Smyth GK (2014) Voom: precision weights unlock linear model analysis
tools for RNA-seq read counts. Genome Biol 15:R29
Lehman MN, Coolen LM, Goodman RL (2010) Minireview: kisspeptin/neurokinin B/dynorphin
(KNDy) cells of the arcuate nucleus: a central node in the control of gonadotropin-releasing
hormone secretion. Endocrinology 151(8):3479–3489
Lehrbach NJ, Armisen J, Lightfoot HL, Murfitt KJ, Bugaut A, Balasubramanian S, Miska EA
(2009) LIN-28 and the poly(U) polymerase PUP-2 regulate let-7 microRNA processing in
Caenorhabditis elegans. Nat Struct Mol Biol 16(10):1016–1020
Li H, Durbin R (2009) Fast and accurate short read alignment with Burrows-Wheeler transform.
Bioinformatics 25:1754–1760
Li H, Ruan J, Durbin R (2008) Mapping short DNA sequencing reads and calling variants using
mapping quality scores. Genome Res 18:1851–1858
Liang K, Keles S (2012) Detecting differential binding of transcription factors with ChIP-seq.
Bioinformatics 28:121–122
Liao Y, Smyth GK, Shi W (2014) Featurecounts: an efficient general purpose program for assigning
sequence reads to genomic features. Bioinformatics 30:923–930
Lomniczi A, Garcia-Rudaz C, Ramakrishnan R, Wilmont B, Khouangsathiene S, Ferguson B,
Dissen GA, Ojeda SR (2012) A single nucleotide polymorphism in the EAP1 gene is associated
with amenorrhea/oligomenorrhea in nonhuman primates. Endocrinology 153(1):339–349
Lomniczi A, Loche A, Castellano JM, Ronnekleiv OK, Bosh M, Kaidar G, Knoll JG, Wright H,
Pfeifer GP, Ojeda SR (2013a) Epigenetic control of female puberty. Nat Neurosci 16:281–289.
[This study describes how the Polycomb family of epigenetic repressors control the epigenetic
landscape at the Kiss1 promoter region, keeping ARC Kiss1 transcription low during the
infantile period]
Lomniczi A, Wright H, Castellano JM, Sonmez K, Ojeda SR (2013b) A system biology approach to
identify regulatory pathways underlying the neuroendocrine control of female puberty in rats
and nonhuman primates. Horm Behav 64(2):175–186
8 Epigenetic and Transcriptional Regulation of the Reproductive Hypothalamus 233
Lomniczi A, Wright H, Castellano JM, Matagne V, Toro CA, Ramaswamy S, Plant TM, Ojeda SR
(2015) Epigenetic regulation of puberty via Zinc finger protein-mediated transcriptional repres-
sion. Nat Commun 6:10195. [This study describes how the GATAD1/LSD1 controls the
epigenetic landscape at the Kiss1 promoter region in rhesus macaques, keeping ARC Kiss1
transcription low during the infantile period]
Maere S, Heymans K, Kuiper M (2005) Bingo: a cytoscape plugin to assess overrepresentation of
gene ontology categories in biological networks. Bioinformatics 21:3448–3449
Maisonpierre PC, Le Beau MM, Espinosa R III, Ip NY, Belluscio L, de la Monte SM, Squinto S,
Furth ME, Yancopoulos GD (1991) Human and rat brain-derived neurotrophic factor and
neurotrophin-3: gene structures, distributions, and chromosomal localizations. Genomics
10:558–568
Margolin AA, Nemenman I, Basso K, Wiggins C, Stolovitzky G, Dalla Favera R, Califano A
(2006) Aracne: an algorithm for the reconstruction of gene regulatory networks in a mammalian
cellular context. BMC Bioinformatics 7(Suppl 1):S7
Mastronardi C, Smiley GG, Raber J, Kusakabe T, Kawaguchi A, Matagne V, Dietzel A, Heger S,
Mungenast AE, Cabrera R, Kimura S, Ojeda SR (2006) Deletion of the Ttf1 gene in
differentiated neurons disrupts female reproduction without impairing basal ganglia function.
J Neurosci 26:13167–13179
Meijs-Roelofs HMA, Kramer P, Cappellen WAV, Woutersen PJA (1985) Changes in serum
concentration and ovarian content of 5a-androstane-3a-17b-diol in the female rat approaching
first ovulation. Biol Reprod 32:301–308
Mortazavi A, Williams BA, McCue K, Schaeffer L, Wold B (2008) Mapping and quantifying
mammalian transcriptomes by RNA-Seq. Nat Methods 5:621–628
Mostafavi S, Ray D, Warde-Farley D, Grouios C, Morris Q (2008) Genemania: a real-time
multiple association network integration algorithm for predicting gene function. Genome Biol
9(Suppl 1):S4
Mueller JK, Dietzel A, Lomniczi A, Loche A, Tefs K, Kiess W, Danne T, Ojeda SR, Heger S (2011)
Transcriptional regulation of the human KiSS1 gene. Mol Cell Endocrinol 342(1–2):8–19
Mueller JK, Koch I, Lomniczi A, Loche A, Rulfs T, Castellano JM, Kiess W, Ojeda S, Heger S
(2012) Transcription of the human EAP1 gene is regulated by upstream components of a
puberty-controlling tumor suppressor gene network. Mol Cell Endocrinol 351(2):184–198
Navarro VM, Gottsch ML, Wu M, Garcia-Galiano D, Hobbs SJ, Bosch MA, Pinilla L, Clifton DK,
Dearth A, Ronnekleiv OK, Braun RE, Palmiter RD, Tena-Sempere M, Alreja M, Steiner RA
(2011) Regulation of NKB pathways and their roles in the control of Kiss1 neurons in the
arcuate nucleus of the male mouse. Endocrinology 152(11):4265–4275
Ojeda SR (1991) The mystery of mammalian puberty: how much more do we know? Perspect Biol
Med 34:365–383
Ojeda SR, Hill J, Hill DF, Costa ME, Tapia V, Cornea A, Ma YJ (1999) The Oct-2 POU-domain
gene in the neuroendocrine brain: a transcriptional regulator of mammalian puberty. Endocri-
nology 140:3774–3789
Ojeda SR, Terasawa E (2002) Neuroendocrine regulation of puberty. In: Pfaff D et al (eds)
Hormones, brain and behavior, vol 4. Elsevier, New York, pp 589–659
Ojeda SR, Skinner MK (2006) Puberty in the rat. In: Neill JD (ed) The physiology of reproduction,
3rd edn. Academic Press/Elsevier, San Diego, pp 2061–2126
Ong KK, Elks CE, Li S, Zhao JH, Luan J, Andersen LB, Bingham SA, Brage S, Smith GD,
Ekelund U, Gillson CJ, Glaser B, Golding J, Hardy R, Khaw KT, Kuh D, Luben R, Marcus M,
McGeehin MA, Ness AR, Northstone K, Ring SM, Rubin C, Sims MA, Song K, Strachan DP,
Vollenweider P, Waeber G, Waterworth DM, Wong A, Deloukas P, Barroso I, Mooser V, Loos
RJ, Wareham NJ (2009) Genetic variation in LIN28B is associated with the timing of puberty.
Nat Genet 41(6):729–733
Perry JR, Stolk L, Franceschini N, Lunetta KL, Zhai G, McArdle PF, Smith AV, Aspelund T,
Bandinelli S, Boerwinkle E, Cherkas L, Eiriksdottir G, Estrada K, Ferrucci L, Folsom AR,
Garcia M, Gudnason V, Hofman A, Karasik D, Kiel DP, Launer LJ, van MJ, Nalls MA,
234 C. F. Aylwin and A. Lomniczi
Rivadeneira F, Shuldiner AR, Singleton A, Soranzo N, Tanaka T, Visser JA, Weedon MN,
Wilson SG, Zhuang V, Streeten EA, Harris TB, Murray A, Spector TD, Demerath EW,
Uitterlinden AG, Murabito JM (2009) Meta-analysis of genome-wide association data identifies
two loci influencing age at menarche. Nat Genet 41(6):648–650
Pierson E, GTEx Consortium, Koller D, Battle A, Mostafavi S, Ardlie KG, Getz G, Wright FA,
Kellis M, Volpi S, Dermitzakis ET (2015) Sharing and specificity of co-expression networks
across 35 human tissues. PLoS Comput Biol 11:E1004220
Pinilla L, Aguilar E, Dieguez C, Millar RP, Tena-Sempere M (2012) Kisspeptins and reproduction:
physiological roles and regulatory mechanisms. Physiol Rev 92(3):1235–1316
Robinson MD, Smyth GK (2007) Moderated statistical tests for assessing differences in tag
abundance. Bioinformatics 23:2881–2887
Robinson MD, Mccarthy DJ, Smyth GK (2010) Edger: a bioconductor package for differential
expression analysis of digital gene expression data. Bioinformatics 26:139–140
Roth CL, Mastronardi C, Lomniczi A, Wright H, Cabrera R, Mungenast AE, Heger S, Jung H,
Dubay C, Ojeda SR (2007) Expression of a tumor-related gene network increases in the
mammalian hypothalamus at the time of female puberty. Endocrinology 148(11):5147–5161.
[This study was the first proposing a tumor-related gene network that controls mammalian
pubertal development]
Sandau US, Mungenast AE, Alderman Z, Sardi SP, Fogel AI, Taylor B, Parent AS, Biederer T,
Corfas G, Ojeda SR (2011a) SynCAM1, a synaptic adhesion molecule, is expressed in
astrocytes and contributes to erbB4 receptor-mediated control of female sexual development.
Endocrinology 152(6):2364–2376
Sandau US, Mungenast AE, McCarthy J, Biederer T, Corfas G, Ojeda SR (2011b) The synaptic cell
adhesion molecule, SynCAM1, mediates astrocyte-to-astrocyte and astrocyte-to-GnRH neuron
adhesiveness in the mouse hypothalamus. Endocrinology 152(6):2353–2363
Sangiao-Alvarellos S, Manfredi-Lozano M, Ruiz-Pino F, Navarro V, Sanchez-Garrido MA,
Leon S, Dieguez C, Cordido F, Matagne V, Dissen GA, Ojeda SR, Pinilla L, Tena-Sempere
M (2013) Changes in hypothalamic expression of the Lin28/let-7 system and related
microRNAs during postnatal maturation and after experimental manipulations of puberty.
Endocrinology 154:942–955
Seminara SB, Messager S, Chatzidaki EE, Thresher RR, Acierno JS Jr, Shagoury JK, Bo-Abbas Y,
Kuohung W, Schwinof KM, Hendrick AG, Zahn D, Dixon J, Kaiser UB, Slaugenhaupt SA,
Gusella JF, O'Rahilly S, Carlton MB, Crowley WF Jr, Aparicio SA, Colledge WH (2003) The
GPR54 gene as a regulator of puberty. N Engl J Med 349:1614–1627
Shao Z, Zhang Y, Yuan GC, Orkin SH, Waxman DJ (2012) Manorm: a robust model for
quantitative comparison of ChIP-seq data sets. Genome Biol 13:R16
Shi Y, Lan F, Matson C, Mulligan P, Whetstine JR, Cole PA, Casero RA, Shi Y (2004) Histone
demethylation mediated by the nuclear amine oxidase homolog LSD1. Cell 119(7):941–953
Shilatifard A (2012) The COMPASS family of histone H3K4 methylases: mechanisms of regula-
tion in development and disease pathogenesis. Annu Rev Biochem 81:65–95
Simon JA, Kingston RE (2009) Mechanisms of polycomb gene silencing: knowns and unknowns.
Nat Rev Mol Cell Biol 10(10):697–708
Steeg PS, Ouatas T, Halverson D, Palmieri D, Salerno M (2003) Metastasis suppressor genes: basic
biology and potential clinical use. Clin Breast Cancer 4:51–62
Steuer R, Kurths J, Daub CO, Weise J, Selbig J (2002) The mutual information: detecting and
evaluating dependencies between variables. Bioinformatics 18(Suppl 2):S231–S240
Sulem P, Gudbjartsson DF, Rafnar T, Holm H, Olafsdottir EJ, Olafsdottir GH, Jonsson T,
Alexandersen P, Feenstra B, Boyd HA, Aben KK, Verbeek AL, Roeleveld N, Jonasdottir A,
Styrkarsdottir U, Steinthorsdottir V, Karason A, Stacey SN, Gudmundsson J, Jakobsdottir M,
Thorleifsson G, Hardarson G, Gulcher J, Kong A, Kiemeney LA, Melbye M, Christiansen C,
Tryggvadottir L, Thorsteinsdottir U, Stefansson K (2009) Genome-wide association study
identifies sequence variants on 6q21 associated with age at menarche. Nat Genet 41(6):734–738
8 Epigenetic and Transcriptional Regulation of the Reproductive Hypothalamus 235
Toro CA, Wright H, Aylwin CF, Ojeda SR, Lomniczi A (2018) Trithorax dependent changes in
chromatin landscape at enhancer and promoter regions drive female puberty. Nat Commun 9
(1):57. [This study describes how the Trithorax family of epigenetic activators control the
epigenetic landscape at the Kiss1 promoter/enhancer regions during the pubertal actiation on
Kiss1 transcription in the ARC]
Urbanski HF, Ojeda SR (1985) The juvenile-peripubertal transition period in the female rat:
establishment of a diurnal pattern of pulsatile luteinizing hormone secretion. Endocrinology
117:644–649
Urbanski HF, Ojeda SR (1987) Gonadal-independent activation of enhanced afternoon luteinizing
hormone release during pubertal development in the female rat. Endocrinology 121:907–913
Wakabayashi Y, Nakada T, Murata K, Ohkura S, Mogi K, Navarro VM, Clifton DK, Mori Y,
Tsukamura H, Maeda K, Steiner RA, Okamura H (2010) Neurokinin B and dynorphin A in
kisspeptin neurons of the arcuate nucleus participate in generation of periodic oscillation of
neural activity driving pulsatile gonadotropin-releasing hormone secretion in the goat. J
Neurosci 30(8):3124–3132
Wang Z, Gerstein M, Snyder M (2009) RNA-seq: a revolutionary tool for transcriptomics. Nat Rev
Genet 10:57–63
Xue Z, Huang K, Cai C, Cai L, Jiang CY, Feng Y, Liu Z, Zeng Q, Cheng L, Sun YE, Liu JY,
Horvath S, Fan G (2013) Genetic programs in human and mouse early embryos revealed by
single-cell RNA sequencing. Nature 500:593–597
Yeung KT, Das S, Zhang J, Lomniczi A, Ojeda S, Xu CF, Neubert TA, Samuels HH (2011) A novel
transcription complex that selectively modulates apoptosis of breast cancer cells through
regulation of FASTKD2. Mol Cell Biol 31(11):2287–2298
Zhang Y, Liu T, Meyer CA, Eeckhoute J, Johnson DS, Bernstein BE, Nusbaum C, Myers RM,
Brown M, Li W, Liu XS (2008) Model-based analysis of ChIP-seq (Macs). Genome Biol 9:
R137
Epigenetic Regulation of the GnRH
and Kiss1 Genes: Developmental 9
Perspectives
Abstract
Gonadotropin-releasing hormone (GnRH) and kisspeptin neurons are indispens-
able for reproductive function. This chapter focuses on the epigenetic regulation
of the GnRH and kisspeptin (Kiss1) genes in the context of neuronal develop-
ment, puberty onset, and maintenance of adult reproductive function. Proper
function of GnRH and kisspeptin neurons in the hypothalamus requires coordi-
nated embryonic and postnatal maturation. Recent studies indicate that diverse
epigenetic phenomena including the formation of chromatin loops, activation of
bivalent domains, maintenance of “stable” histone modifications, and DNA
methylation, active demethylation, and hydroxymethylation are all involved in
epigenetic regulation of these neurons. This chapter focuses on how these epige-
netic regulation components interact with each other and which enzymes or
binding factors are involved in specific stages of neuronal development or
aging. By emphasizing similar epigenetic mechanisms operating within these
two peptidergic neuronal populations, we highlight likely targets for future
investigations of environmental influences over puberty timing and progression.
Keywords
Gonadotropin-releasing hormone (GnRH) · Kisspeptin · Puberty · Bivalent
chromatin domain · DNA hydroxymethylation · Tet enzymes
J. R. Kurian
Departments of Obstetrics/Gynecology and Internal Medicine, Southern Illinois University School
of Medicine, Springfield, IL, USA
St. Johns Hospital, Carol Jo Vecchie Women and Children’s Center, Springfield, IL, USA
E. Terasawa (*)
Department of Pediatrics, University of Wisconsin-Madison, Madison, WI, USA
Wisconsin National Primate Research Center, University of Wisconsin-Madison, Madison, WI,
USA
e-mail: [email protected]
The genetic control of GnRH gene expression is well characterized and depends on
several cis-regulatory sequences in the 50 region of the gene (Box 9.1). The rat gene
contains a neuron-specific enhancer region between 1863 and 1571 (Kepa et al.
1992, 1996b; Clark and Mellon 1995; Whyte et al. 1995). This spans a major region
of homology between rat (1786 to 1559) and human (2766 to 2539) genes.
Interestingly, however, this portion of the human gene does not appear to activate
gene expression. In fact, based on serial truncations of the 50 human GnRH gene in
luciferase assay constructs, it appears that this region represses gene expression
9 Epigenetic Regulation of the GnRH and Kiss1 Genes: Developmental Perspectives 239
(Kepa et al. 1996a). Importantly, this area has sequence similarity to a 50 portion of
the rhesus monkey GnRH gene. We noticed that this distal 50 region of the rhesus
monkey gene contains a 243-bp segment (2126 to 1863) that has 60% GC
content and a CpG (cytosine–guanine) dinucleotide observed to expected ratio of
0.65 (14 CpG sites). These characteristics define the region as a CpG island (CGI;
Gardiner-Garden and Frommer 1987). CGIs, when associated with gene promoters,
are related to the epigenetic regulation (DNA methylation) of gene expression
(Deaton and Bird 2011).
Box 9.1 Genetic and Epigenetic Regulation of the GnRH Gene During
Development
Understanding transcriptional regulation of the GnRH gene is important, as
deficiency of this process leads to abnormal timing of puberty and infertility.
The mammalian form of GnRH (GnRH1) encoded by the GnRH gene,
consists of a ten amino acid-peptide (pyro-Glu-His-Trp-Ser-Tyr-Gly-Leu-Arg-
Pro-Gly-NH2), cleaved from the 92-amino acid GnRH precursor protein
(prepro-GnRH peptide). The GnRH gene contains four exons and three
introns. The 50 GnRH gene-regulatory region comprises a proximal promoter
and three enhancer regions. Transcription factors bind to conserved enhancer
regions and the proximal promoter region.
Genetic Regulation: Since the 1990s several groups reported using in vitro
GnRH cell lines and in vivo transgenic mouse approaches to identify various
transcription factors that bind the 50 GnRH gene-regulatory region, and activate
or repress GnRH gene transcription. Among the activators are the homeodomain
transcription factors OCT1, OTX2, Six3/6, MSX1 and VAX1, neuron-specific
elements, GATA4 (Radovick et al. 1994; Clark and Mellon 1995; Lawson et al.
1998; Kelley et al. 2000; Kim et al. 2007; Larder and Mellon. 2009; Gan et al.
2012; Xie et al. 2015; Larder et al. 2013; Hoffmann et al. 2018), an estrogen
response element (human/primates only, Radovick et al. 1991b, 1994), and a
kisspeptin response element (Novaira et al. 2012, 2016). Repressors are C/EBPβ
(Belsham and Mellon 2000; Messina et al. 2016), MSX1/2 (Givens et al. 2005,
Rave-Harel et al. 2005), NKX2.1/TTF-1 (Provenzano et al. 2010), and ZEB1
(Messina et al. 2016). See further discussion in Hoffmann and Mellon (2018).
Epigenetic Regulation: More recently, GnRH gene expression was also
found to be altered by environmental factors through mechanisms such as
DNA methylation and histone modifications without changes in nucleotide
sequences. Recent findings are summarized in this chapter.
Fig. 9.1 Embryonic development of primate GnRH neurons is marked by a dramatic shift in Gnrh
gene expression and active demethylation of a 50 CpG island. (a) Total RNA was extracted from
nasal placode cultures containing GnRH neurons at 0, 14, and 20 days (div). Gnrh mRNA levels,
measured by quantitative PCR, started to increase after 14 div, reaching the highest level at 20 div.
(P < 0.05 versus 0 div; #P ¼ 0.05 versus 14 div). (b) A schematic representation of the rhesus
monkey GnRH gene depicting the location of the 50 CpG island (CGI) and nucleotide sequence of
this region. This region is the only classifiable CGI within 2500 bases upstream of the Gnrh
transcription start site. CpG sites are indicated with superscript numbers corresponding to the CpG
sites in the bottom graph. (c) Nasal placode regions, containing GnRH neurons, were dissected from
two rhesus monkey embryos at E36 and E37, plated and then harvested on 0, 14, or 20 div. DNA
extracted from pooled samples (four cultures) at each time point was bisulfite sequenced. Percent
changes in methylation at each CpG site on 0 div (black bars), 14 div (gray bars), and 20 div (white
bars) are shown. CpG methylation status was significantly higher at 0 div compared with 20 div at
sites 4, 5, 6, 8, 9, 11, 12, and 14 (P 0.01). CpG methylation status was significantly higher at 0 div
compared with 14 div at site 5 (#P < 0.05) and at 0 div compared with 14 div but not 20 div at site
10 (##P < 0.05). Modified from Kurian et al. Endocrinology 151: 5359–5368 (2010a)
242 J. R. Kurian and E. Terasawa
puberty onset. Alternatively, the CGI may become re-methylated during juvenile
development, and again demethylated at puberty onset. The latter scenario would
suggest that lower CpG methylation status must be maintained to enable elevated
GnRH gene expression. Consequently, a mechanism responsible for DNA demeth-
ylation and perhaps another mechanism responsible for maintaining hypomethylated
DNA might be necessary for the transition to puberty and maintenance of reproduc-
tive function. A comparison of CpG methylation status between perinatal and early
pubertal MBH tissue will be necessary to differentiate between these two scenarios.
Nonetheless, our current findings suggest that DNA demethylation is an important
aspect of GnRH neuronal development and function.
DNA demethylation is achieved via two general mechanisms, passive or active.
Passive demethylation occurs through cell divisions and interruption of maintenance
methyltransferase activity. Current pharmacological approaches for DNA demethyl-
ation target this mechanism. For example, 5-azacytidine (5-aza) is metabolized and
subsequently incorporated into DNA, where it then acts as a substrate that covalently
traps DNMTs after methyl transfer to the 5-aza nitrogen (Schermelleh et al. 2005;
Svedruzic 2008). Given the postmitotic/nondividing state of GnRH neurons in our
in vitro cultures and across the pubertal transition measured in MBH samples, the
process of DNA demethylation we discovered must be active. The mechanisms
responsible for active DNA demethylation are not well characterized, and until
recently skepticism has remained over the existence of this process in mammalian
systems. Several mechanisms are now proposed, and intense efforts to validate or
better characterize these mechanisms continue. For a thorough background, we
suggest a recent review (Wu and Zhang 2010) that outlines several promising
pathways discovered during the past decade. Presently, a well-accepted mechanism
is the sequential enzymatic process beginning with the oxidation of
5-methylcytosine (5mC) to 5-hydroxymethylcytosine (5hmC), which is carried out
by anyone of three ten-eleven-translocation enzymes (Tet 1–3). 5hmC is a stable
epigenetic modification, though under certain circumstances it is recognized and
excised by thymine DNA glycosylases. Base excision repair subsequently completes
the transition. Importantly, brain tissue, and particularly the hypothalamus, shows
the highest reported levels of 5-hydroxymethylcytosine (5hmC) (Branco et al. 2012).
Emerging physiological evidence also supports a role for Tet enzymes in neuro-
endocrine development, and particularly the control of reproductive function. For
example, Tet1 knockout mice exhibits deficits in fecundity. This effect is primarily
a female-specific abnormality, as typical litter sizes result from crossing wild-type
females with Tet1 knockout males, whereas wild-type males mated to Tet1 knockout
females produce significantly fewer offspring per litter (Dawlaty et al. 2011). A
subsequent report suggests this defect in fecundity might be the consequence of
abnormal progression of female germ cell development through the second meiotic
division prior to ovulation (Yamaguchi et al. 2012). Interestingly, stimulation of the
second meiotic division and ovulation requires a GnRH-driven LH surge
(Mehlmann 2005). Consequently, given the apparent active DNA demethylation
associated with GnRH neuronal development, the effect of Tet depletion on abnor-
mal reproductive function might be the consequence of altered Tet-mediated
9 Epigenetic Regulation of the GnRH and Kiss1 Genes: Developmental Perspectives 243
Histones are an integral component of nucleosomes, the primary units for genome
organization or compaction. Consequently, through their interaction with DNA,
these proteins have a critical role in determining gene expression patterns. There
are four primary classes of histones, 1 through 4. Histones 2 through 4 are
components of the core octamer, which DNA circumnavigates in about 146 base
pairs to form a single nucleosome. Histone 1 is a scaffold protein, which tightly
packages nucleosomes when present. Several variants of this histone are distinguish-
able by sensitivity to hormones (Banks et al. 2001). Histone 2 also has several
variants including A, AX, and B. Histone 2AX is particularly intriguing in the
context of neuronal maturation (Lee et al. 2010) and function based on its association
with activity-dependent DNA double-strand breaks in neurons (Suberbielle et al.
2013). Histones 3 and 4 complete the nucleosome octamer with histone 3 the most
heavily investigated in the context of neuroendocrine function. Histone 3 has two
variants, H3.3A and H3.3B. These variants are integral to DNA replication-
independent histone switching, which would be presumed an important mechanism
in the regulation of post-mitotic cell activity. To our knowledge, nothing is reported
regarding the relationship between histone switching and hypothalamic neuronal
maturation or function.
Histone proteins package DNA largely due to their predominant positive charge
attracting negatively charged DNA. Posttranslational modifications (PTMs) alter the
strength of that attraction, and recruit or repel transcriptional machinery and histone-
modifying enzymes. Consequently, these PTMs alter gene accessibility and rates of
transcription. Several known modifications include acetylation, phosphorylation,
methylation, ubiquitination, sumoylation, and GlcNAcylation. To date,
measurements of histone PTMs in neuroendocrine systems have focused on acety-
lation and methylation. Acetylation leaves a more negative charge on histones and
consequently promotes transcription. Methylation is neutral, and depending on the
location and degree (mono, di, or tri-methylation), it can either promote or inhibit
transcription.
Mellon and colleagues were the first to report a pattern of permissive histone
modifications enabling elevated or mature Gnrh gene transcription (Iyer et al. 2011).
Their studies capitalized on the distinct stages of development between two GnRH
neuronal cell lines. GN11 cells, originally isolated from a tumor in the mouse nasal
placode (Radovick et al. 1991a), express Gnrh at very low levels, whereas GT1 cells,
244 J. R. Kurian and E. Terasawa
which were isolated from an analogous tumor in the mouse hypothalamus (Mellon
et al. 1990), are characterized by mature activity patterns including elevated Gnrh
expression. In essence, comparisons between these two cell lines are similar to our
evaluations of embryonic nasal placode derived neurons from days 0 and 20 in vitro
described above. They found that the GnRH gene promoter and enhancer regions in
immature GN cells were more heavily associated with a repressive histone modifi-
cation: histone 3 (H3) lysine 9 (K9) di-methylation (me2). On the other hand, the
same genomic regions in GT1 cells were associated with the permissive H3K9
acetylation and H3K4me3 PTMs. For comparison, they also evaluated these histone
PTM patterns in a nonneuronal (NIH3T3) cell line. The repressive PTMs were high
in NIH3T3 cells, intermediate in GN cells and low in GT1 cells. The presence of
permissive PTMs was low and similar between NIH3T3 and GN cells but signifi-
cantly higher in mature GT1 cells.
The intermediate repressive chromatin state in GN cells is indicative of a bivalent
promoter, where both repressive and permissive histone PTMs maintain genes in a
repressed albeit primed position. These chromatin domains were first described in
embryonic stem (ES) cells and associated with developmentally regulated transcrip-
tion factors (Azuara et al. 2006; Bernstein et al. 2006; Pan et al. 2007). Several recent
reports, taken together, indicate that establishment of bivalent domains through
polycomb repressive complex 2 (PRC2) and COMPASS families is critical for
neural lineage differentiation from ES cells (Yu et al. 1995; Yagi et al. 1998; Glaser
et al. 2006; Pasini et al. 2007; Shen et al. 2008). The PRC2 component, Ezh2, is
responsible for H3K27 methylation, which is subsequently bound by the PRC1
complex to maintain tri-methylation at H3K27. COMPASS complexes methylate
H3K4. Together, these complexes are proposed to establish a bivalent promoter with
the heavily repressive H3K27me3 mark in close proximity to H3K4me3. Interest-
ingly, this bivalent domain also appears related to the juvenile repression and
subsequent activation of Kiss1 (kisspeptin) gene expression during the pubertal
transition in female rats (Lomniczi et al. 2013). As a major stimulant of GnRH
release, this epigenetic regulation of Kiss1 expression (discussed in detail below)
may be instrumental in pubertal development. Together, this also suggests that
bivalent domain activation may be a consistent mechanism across reproductive
neuroendocrine systems.
Fig. 9.3 Tet2 influences accumulation and maintenance of the activating H3K4 trimethylation in
immature and differentiated GnRH neurons, respectively. (a) Immature GnRH neuronal cell lines
(GN11) express Gnrh at very low levels and less Tet2 than differentiated GnRH neuronal cell lines
(GT1–7). Increasing Tet2 expression in GN11 cells by transient transfection leads to an elevation of
Gnrh mRNA and H3K4me3 at the Gnrh gene neuron-specific enhancer and promoter. The histone
modifications are suspected to be generated by the Tet2 interacting histone-modifying enzyme (MLL:
mixed lineage leukemia) or complex (SET1: a histone methyltransferase complex) (b) Knockdown of
Tet2 in differentiated GnRH neuronal cultures (GT1–7 cells) leads to a loss of Tet2 binding near the
Gnrh gene promoter, loss of H3K4me3 abundance in the same region and reduction in Gnrh mRNA
levels. Modified from Kurian et al. Endocrinology 157: 3588–3603 (2016)
these mice exhibited a gradual decline in GnRH neurons, from typical numbers and
distribution at P0 to no detectable Gnrh expression in adult animals. These animals
also developed hypogonadotropic hypogonadism and had significantly lower
circulating LH and FSH. While mRNA and miRNA expression patterns have not
been profiled in mice lacking Dicer in GnRH neurons, it is clear that miR-200 and
miR-155, which target the GnRH gene repressors Zeb-1 and Cebpp, respectively, are
critical for GnRH neuron maintenance in rodents (Messina et al. 2016). Moreover,
the work by Messina et al. (2016) supports the concept that a set of microRNAs
regulates Gnrh gene expression during sexual development. Whether additional
miRNAs are involved in rodents, whether a similar mechanism can be seen in
primates, and whether the microRNA-dependent epigenetic regulation of GnRH
release is involved in the pathophysiology of disease states, such as idiopathic
hypogoanadotropic hypogonadism in human patients remain open questions.
248 J. R. Kurian and E. Terasawa
Kisspeptin peptides were originally coined “metastins” after they were discovered in
the mid-1990s as novel ligands that suppressed metastasis of breast cancer and
melanoma (Lee et al. 1996). Since that discovery, the biological roles of kisspeptin
have broadened, most notably for their participation in reproductive maturation
(puberty) and cyclic ovulation. In 2003, two independent reports revealed a novel
signaling mechanism crucial to reproductive function. Specifically, mutations in the
human gene encoding G-protein coupled receptor 54 (GPR54 or KISS1R) were
linked to the absence of puberty or severe disruption of pubertal progression
(Seminara et al. 2003; de Roux et al. 2003). More recently, a familial mutation in
KISS1 (the gene encoding kisspeptins) was linked to the impairment of pubertal
progress (Topaloglu et al. 2012). Mice lacking either Gpr54 or Kiss1 exhibits
delayed puberty and hypogonadotropic hypogonadism (Funes et al. 2003;
d’Anglemont de Tassigny et al. 2007; Lapatto et al. 2007), whereas exogenous
kisspeptin exposures potently stimulate gonadotropin release in prepubertal rodents,
sheep, and primates (Gottsch et al. 2004; Messager et al. 2005; Navarro et al.
2005a, b; Shahab et al. 2005; Dhillo et al. 2007; Guerriero et al. 2012). These
kisspeptin stimulations of gonadotropin release occur through activation of GPR54
(KISS1R) located on GnRH neurons (Hrabovszky et al. 2010; Kirilov et al. 2013;
Yeo et al. 2014), suggesting GnRH neurons are capable of adult levels of activity
prior to puberty onset, and that developmental activation of kisspeptin systems may
play a role in the timing of puberty. Consequently, based on the proposed environ-
mental influence over timing of puberty, significant efforts are currently directed
toward characterizing the epigenetic development and regulation of kisspeptin
neurons.
Kisspeptin cell bodies are concentrated in the arcuate nucleus (ARC) and the
anteroventral periventricular nucleus (AVPV)/periventricular nucleus (PEN)
(Gottsch et al. 2004; Smith et al. 2005; Clarkson and Herbison 2006; Kauffman
et al. 2007). In rodents, the presence of kisspeptin cell bodies in the medial amygdala
is also described (Kim et al. 2011). While KISS1 expression in the ARC is consistent
across species (Smith et al. 2005, 2010; Ramaswamy et al. 2008; Shahab et al. 2005;
Smith 2008; Alcin et al. 2013), there are some species differences in the distribution
pattern of kisspeptin neurons in the anterior population. In rodents, the AVPV
population is very clear but in primates, including humans, a similar but less
conspicuous population is present (Rometo et al. 2007; Hrabovszky et al. 2012;
Smith et al. 2010; Watanabe et al. 2014; Vargas Trujillo et al. 2017). In sheep and
goats, in contrast, kisspeptin neurons are diffusely distributed in the preoptic area
(POA), although within the POA a surge activated population is restricted in the
AVPV area (Franceschini et al. 2006; Smith et al. 2009; Hoffman et al. 2011;
Matsuda et al. 2015). A clear sexual dimorphism of the AVPV population, i.e., a
larger number of kisspeptin neurons in female, exists in mice and rats (Clarkson and
Herbison 2006; Kauffman et al. 2007; Ohkura et al. 2009, Semaan and Kauffman
2010), and this reflects the significance of AVPV kisspeptin neurons in the preovu-
latory surge in rodents. However, there is little sex difference in the number of the
9 Epigenetic Regulation of the GnRH and Kiss1 Genes: Developmental Perspectives 249
anterior kisspeptin neurons in sheep and monkeys and anterior kisspeptin neurons
respond to the estrogen challenge in castrated male monkeys, but not in the castrated
ram (Cheng et al. 2010; Watanabe et al. 2014). Therefore, despite the well-established
concept that negative feedback effects of estradiol are mediated through estrogen
receptor alpha (ERα) located in kisspeptin neurons in the ARC, whereas the positive
feedback effects of estradiol are mediated through ERα in kisspeptin neurons in the
rodent AVPV (Smith 2013; Herbison 2016), the role of kisspeptin neurons in
regulation of GnRH release in primates and ruminants requires further investigation.
In rodents, kisspeptin neuronal systems develop in characteristic regional and
sex-specific patterns. Kiss1 is detectable in the rat and mouse ARC by E13–14
(Desroziers et al. 2012a; Kumar et al. 2014) and persists stably throughout postnatal
and pubertal development (Poling and Kauffman 2013; Semaan and Kauffman
2015). In contrast, AVPV Kiss1 expression becomes apparent at P7–10 (Semaan
et al. 2010; Desroziers et al. 2012b; Clarkson et al. 2009) with only a small amount
of kisspeptin peptide detectable by P15 (Clarkson et al. 2009) but levels substantially
elevate prior to the onset of puberty (Clarkson et al. 2009; Semaan and Kauffman
2015). In addition, the appearance of Kiss1 AVPV briefly lags behind in males and is
significantly lower in pubertal and adult males compared to females (Kauffman et al.
2007; Semaan et al. 2010; Clarkson and Herbison 2006; Homma et al. 2009). The
sex- and region-specific differences in kisspeptin expression are critical to
sex-specific reproductive function and have therefore been a focus of early
investigations of rodent kisspeptin neuron epigenetic regulation. The origin and
prenatal development of prenatal nonhuman primate kisspeptin neurons have not
been studied. Nevertheless, the kisspeptin neurons during the neonatal mini-puberty
are active at levels consistent with adult neurons, and subsequently enter a dormant
state during the prepubertal period until the onset of puberty (Ramaswamy et al.
2013; Shahab et al. 2018).
Sex differences in the brain are largely created during a prenatal or perinatal “critical
period” by a sex-specific steroid environment (Phoenix et al. 1959). During this
critical period, a male-specific acute surge of testicular testosterone secretion
coordinates the masculinization/defeminization of the brain and associated physiol-
ogy and behaviors (Simerly 1998, 2002). Importantly, prior to acute sex steroid
exposure, the brain exhibits the potential to develop both male- and female-specific
neural circuitry. While in rodents testosterone aromatized into estradiol is responsi-
ble for this organizational action (Arnold and Gorski 1984; McCarthy 1994), in
ruminants and primates, androgens themselves are responsible (Wallen 2005;
Puttabyatappa and Padmanabhan 2017). As evidence, male-like brain development
can be initiated by perinatal estradiol treatment of female rats and mice, whereas
androgen treatment of early to mid-embryonic sheep and monkey embryos directs
male-like brain development (Phoenix et al. 1959; Arai and Gorski 1968; see Wallen
250 J. R. Kurian and E. Terasawa
2009). Conversely, castration of newborn male rats and mice leads to a female-like
brain in adulthood (Döhler et al. 1984). The critical period in rodents extends
through approximately P10, after which sex steroid manipulations no longer alter
sexually dimorphic brain structures (Gorski 1985). Some brain traits, however,
appear to be influenced by steroid exposures during the pubertal stage (Schulz and
Sisk 2006; Schulz et al. 2009).
Importantly, in addition to sex steroids, sexually dimorphic experiences during
the perinatal critical period also appear to masculinize the brain. For example, male
rodents experience a higher degree of maternal grooming than female littermates
during this period. Subjecting female rats to a sex steroid surge or simulated maternal
grooming in the first week of life is reported to masculinize the ERα gene CpG
methylation pattern and gene expression profile in the MBH (Kurian et al. 2010b).
That study was the first to report an equivalent impact of steroids and environmental
influence over sexual differentiation of epigenetic patterns in the brain. This
suggested that epigenetic mechanisms likely had a significant role in the sexual
differentiation of the brain in general. In fact, several “critical periods” for sex steroid
exposures were soon after related to epigenetic factors, including bed nucleus of the
stria terminalus (BNST) size and vasopressin fiber projections (Murray et al. 2011),
male sexual behavior in rats and olfactory behavior in mice (Matsuda et al. 2011),
and steroid receptor and DNA methyltransferase 3a expression (Schwarz et al. 2010;
Kolodkin and Auger 2011).
To date, only one study has investigated whether sexual differences of the two
kisspeptin systems are driven by epigenetic factors. Specifically, Semaan,
Kauffman, and colleagues evaluated whether histone deacetylation and DNA meth-
ylation contributed to sexually dimorphic regulation of AVPV Kiss1 gene expres-
sion and cell number (Semaan et al. 2012). They pharmacologically blocked histone
deacetylation during the postnatal critical period with valproic acid, and evaluated
male and female AVPV Kiss1 expression in adulthood. While the authors observed
the expected disruption of sex differences in the BNST similar to that reported by
Murray et al. (2011), Kiss1 gene expression in the AVPV was elevated in both sexes
compared to same sex controls and there was no sex-specific effect per se. Whether,
the elevation of Kiss1 expression in males had a functional impact (i.e., feminization
of estradiol-induced LH surge potential) was not evaluated and remains an important
question. Additionally, potential off-target effects of valproic acid are a significant
concern in interpreting these results, as valproic acid is also a voltage-gated sodium
channel blocker and elevates GABA levels in the brain, which could indirectly alter
GnRH neuronal activity.
In the same study, Semaan et al. (2012) further evaluated DNA methylation over
70 unique sites of the Kiss1 gene in microdissected AVPV and ARC tissues. CpG
methylation was significantly higher in females, primarily at the putative Kiss1
promoter in the AVPV but not in the ARC. This finding was unexpected, as
methylation is typically associated with reduced gene expression, though AVPV
Kiss1 expression is substantially higher in females than males. Consequently, it is
currently unclear how methylation alters Kiss 1 gene expression, though it is likely
involved in chromatin conformation, which is critical to hormone-dependent Kiss
9 Epigenetic Regulation of the GnRH and Kiss1 Genes: Developmental Perspectives 251
1 expression in the AVPV (discussed in the following section). Whether the perinatal
hormone surge drives epigenetic changes in the Kiss1 gene in the AVPV also
remains an open critical question.
Fig. 9.4 Estradiol (E2) exposure has brain region-specific influence over Kiss1 gene expression,
chromatin conformation and histone 3 (H3) acetylation (Tomikawa et al. 2012). (a) A schematic
representation of the mouse Kiss1 gene depicts several critical regions as defined by color and shapes
including a 50 region (black), a coding region (green) that includes a promoter (lime green box) 3 exons
(numbered boxes) and a CpG island (red box, within the third exon), a 30 region (blue) and a
transcriptional modifier region (red). Small gray shapes indicate sites utilized for analysis by
Tomikawa et al. (2012). The table indicates the direction of H3 acetylation abundance detected in
each gene region in the ARC and AVPV during E2 exposure. H3 acetylation significantly increased
("), decreased (#), or remained the same ($); N.D. not determined. (b) Representations of DNA
looping at the Kiss1 gene in the ARC and AVPV. Yellow circles represent H3 acetylation abundance.
Location of yellow circles enables visualization of Kiss1 regional interactions with the acetylated
histone. Prior to E2 exposure, the Kiss1 gene structure is similar in the ACR and AVPV, though a
second loop is apparent in only the ARC. E2 exposure causes formation of a complex DNA structure
and decrease in Kiss1 expression in the ARC. On the other hand, the AVPV 50 loop is lost in favor of a
long range 30 loop and increased Kiss1 expression during E2 exposure. Created, based on Tomikawa
et al., Proc Natl Acad Sci USA 109:E1294-1301 (2012), Uenoyama et al., Neuroendocrinology
103:640–649 (2016), and Kurian et al., Endocrinology 157: 3588–3603 (2016)
Kisspeptin signaling is essential for puberty onset and progression. Because puberty
onset is heavily influenced by environmental conditions including climate, stress,
body weight, and environmental estrogens, interest in the epigenetic regulation of
kisspeptin cell development and puberty-related Kiss1 gene expression has grown
substantially. As mentioned earlier, similar to the findings of Mellon and colleagues
comparing Gnrh gene promoter structure of immature and mature cell lines,
Lomniczi et al. (2013) reported that the female rat pubertal transition is accompanied
by increased prevalence of activating histone PTMs at the Kiss1 promoter. Specifi-
cally, prior to puberty onset, the kisspeptin promoter was associated with a bivalent
domain (i.e., H3K27me3 and H3K4me3) and occupied by a component of a
polycomb repressive complex, EED. The transition to puberty from prepubertal
period was accompanied by decreased EED occupancy of the kisspeptin promoter,
a gradual loss of the repressive H3K27me3 PTM, and increased levels of the
permissive H3K9,14 acetylation and H3K4me3 PTMs. The authors suggest this
process is the consequence of increased DNA methylation of the EED promoter,
leading to lower EED expression and consequential decreased occupancy of the
kisspeptin promoter by the repressive PRC2 (EED, SUZ12, Ezh2) complex. How-
ever, these conclusions were largely based on observations of dramatically delayed
puberty in female mice when DNA methylation was inhibited by peripheral admin-
istration of the pharmacological DNA methyltransferase inhibitor 5-azacytidine
(5-aza). While the findings are consistent with the expected results, this 5-aza
activity may not be specific to inhibition of postmitotic neuronal DNA methylation.
Because the well-characterized mechanism of 5-aza requires nucleoside
incorporation into DNA (Schermelleh et al. 2005; Stresemann and Lyko 2008;
Svedruzic 2008) the mechanism by which this compound inhibits DNMT in
postmitotic cells remains unclear. In addition, the significant reduction in growth
254 J. R. Kurian and E. Terasawa
rate and elevated levels of circulating corticosterone that occurs following initiation
of drug treatment (Lomniczi et al. 2013) is indicative of toxicity that likely
contributes to delays in sexual maturation. Because of these concerns, more direct
approaches (e.g., cell-specific genetic or enzyme expression manipulations) are
necessary to clarify the relationship between DNA methylation, chromatin
modifications, and puberty onset. Nonetheless, these studies and a recent report
that isolated another vital repressor gene (GATAD1) through an extensive gene
array analysis in orchiectomized male monkey hypothalamus (Lomniczi et al. 2015),
further establish the concept that structural modification of bivalent promoters in the
neuroendocrine hypothalamus is an integral step toward puberty onset and repro-
ductive function in primates and rodents.
Lomniczi and colleagues’ focused approach toward measuring histone modifica-
tion status during development at one gene has tremendous value for clarifying the
temporal progression of bivalent promoter transactivation. Importantly, they report
that permissive histone PTMs (H3K4me3 and H3K9,14 acetylation) accumulate
near the kisspeptin promoter during the transition from juvenile to early pubertal
stages. This preceded the loss of the repressive H3K27me3 PTM, suggesting that the
recruitment of activating complexes is imperative for transactivation and increased
gene expression. Based on the temporal relationship between H3K4me3 accumula-
tion and loss of H3K27me3, as well as the association between Tet2 and H3K4me3,
we recently evaluated the impact of kisspeptin neuron-specific ablation of Tet2
(kTKO animals) on puberty onset and progression in mice (Kurian et al. 2014).
Remarkably, Tet2 disruption delayed puberty onset and progression in both male
and female kTKO animals. Whether this impact on puberty is the consequence of
Tet2-mediated DNA demethylation, altered DNA hydroxymethylation patterns,
H3K4me3 accumulation or another mechanism remain open and important
questions. More recent evidence indicates that histone-modifying Tet2 interactors
(i.e., MLL1 and MLL3) are also critical mediators of increased rodent kiss1 expres-
sion and pubertal progression; suggesting factors that drive Tet2/MLL interactions
could represent signals that are critical for initiating puberty (Toro, et al. 2018).
Which genomic regions in kisspeptin neurons, in addition to Kiss1, are altered
following Tet2 disruption also remain unknown.
The similar genomic targets as integral positions of epigenetic regulation in
kisspeptin cells are predicated on the recent discoveries of additional peptides that
are co-expressed in ARC kisspeptin neurons, namely neurokinin B (Tac2) and
dynorphin. Neurons that express all three peptides are known as KNDy neurons
(Goodman et al. 2014). In fact, Tac2 expression in the ARC in mice is an early
marker of puberty onset (Gill et al. 2012a), and is likely regulated by a bivalent
promoter. As evidence, recent preliminary studies (Gill et al. 2012b) found that
female mice heterozygous for a loss-of-function mutation in Lsd1, a histone lysine
demethylase of di- and mono-methylated H3K4, exhibit precocious vaginal opening
and first ovulation (3–4 days prior to wild-type littermates), with early elevations of
circulating gonadotropins and hypothalamic expression of Tac2 (Gill et al. 2012b).
Importantly, these findings parallel those of Lomniczi et al. (2013), in that lower
expression of an epigenetic repressive enzyme is related to the elevated
9 Epigenetic Regulation of the GnRH and Kiss1 Genes: Developmental Perspectives 255
9.9 Conclusion
Taken together, the findings detailed in this chapter highlights exciting connections
between epigenetic regulations in different cells of the reproductive neuroendocrine
hypothalamus. Nevertheless, our understanding of epigenetic regulation of GnRH
and kisspeptin neurons and reproductive neuroendocrine function, in general, is still
in its infancy. In fact, the findings described in this chapter are all based on changes
in known epigenetic markers in a given physiological condition and the question of
whether actual modifications of environmental conditions result in similar epigenetic
modifications in GnRH and kisspeptin neurons, along with altered physiology, has
not been examined to date. For example, it is known that exposure of prepubertal
animals to stress results in delayed puberty, whereas exposure of prepubertal animals
to a high-calorie diet induces precocious puberty. Do these environmental changes
cause consistent opposite epigenetic modifications to GnRH? The field is
progressing rapidly. We are inching closer to the ultimate promise of neuroendocrine
epigenetic research: to define how the environment influences our physiology.
256 J. R. Kurian and E. Terasawa
Acknowledgments The authors express their sincere appreciation for many current and past
postdoctoral research fellows, graduate students, and research specialists including Kim L. Keen,
who has contributed to research in the Terasawa laboratory as well as Somaja Louis and Jennifer Li,
who contributed to research in the Kurian laboratory. This work was supported by the NIH grants,
HD011355, HD015344, and HD077447 for ET, ES020878 for JRK, and OD01106 and RR0061 for
the WNPRC.
Key References
Clark and Mellon (1995). This is the first paper identifying that the homeodomain
transcription factor Oct 1 regulates the GnRH gene transcription.
Iyer et al. (2011). This is the first paper demonstrating chromatin modifications of
GnRH gene expression during neuronal differentiation.
Kepa et al. (1996a). This is one of the first works to analyze the structure of GnRH
gene promoter.
Kurian et al. (2010a). This is the first paper to demonstrate that epigenetic changes
occur during GnRH neuronal maturation.
Lomniczi et al. (2015). This is a seminal work demonstrating that Zinc finger protein
is involved in prepubertal GnRH suppression in non-human primates.
Messina et al. (2016). This paper identified that microRNAs regulate GnRH gene
expression, which correlates with the timing of puberty.
References
Alcin E, Sahu A, Ramaswamy S, Hutz ED, Keen KL, Terasawa E, Bethea CL, Plant TM (2013)
Ovarian regulation of kisspeptin neurones in the arcuate nucleus of the rhesus monkey (Macaca
mulatta). J Neuroendocrinol 5:488–496
Arai Y, Gorski RA (1968) Critical exposure time for androgenization of the developing hypothala-
mus in the female rat. Endocrinology 82:1010–1014
Arnold AP, Gorski RA (1984) Gonadal steroid induction of structural sex differences in the central
nervous system. Annu Rev Neurosci 7:413–442
Azuara V, Perry P, Sauer S, Spivakov M, Jørgensen HF, John RM, Gouti M, Casanova M,
Warnes G, Merkenschlager M, Fisher AG (2006) Chromatin signatures of pluripotent cell
lines. Nat Cell Biol 8:532–538
Banks GC, Deterding LJ, Tomer KB, Archer TK (2001) Hormone-mediated dephosphorylation of
specific histone H1 isoforms. J Biol Chem 276:36467–36473
Belsham DD, Mellon PL (2000) Transcription factors Oct-1 and C/EBPbeta (CCAAT/enhancer-
binding protein-beta) are involved in the glutamate/nitric oxide/cyclic-guanosine 5'-
monophosphate-mediated repression of mediated repression of gonadotropin-releasing hor-
mone gene expression. Mol Endocrinol 14:212–228
Bernstein BE, Mikkelsen TS, Xie X, Kamal M, Huebert DJ, Cuff J, Fry B, Meissner A, Wernig M,
Plath K, Jaenisch R, Wagschal A, Feil R, Schreiber SL, Lander ES (2006) A bivalent chromatin
structure marks key developmental genes in embryonic stem cells. Cell 125:315–326
Bestor TH, Ingram VM (1983) Two DNA methyltransferases from murine erythroleukemia cells:
purification, sequence specificity, and mode of interaction with DNA. Proc Natl Acad Sci U S A
80:5559–5563
9 Epigenetic Regulation of the GnRH and Kiss1 Genes: Developmental Perspectives 257
Bestor T, Laudano A, Mattaliano R, Ingram V (1988) Cloning and sequencing of a cDNA encoding
DNA methyltransferase of mouse cells. The carboxyl-terminal domain of the mammalian
enzymes is related to bacterial restriction methyltransferases. J Mol Biol 203:971–983
Branco MR, Ficz G, Reik W (2012) Uncovering the role of 5-hydroxymethylcytosine in the
epigenome. Nat Rev Genet 13:7–13
Carey BW, Finley LW, Cross JR, Allis CD, Thompson CB (2015) Intracellular α-ketoglutarate
maintains the pluripotency of embryonic stem cells. Nature 518(7539):413–416
Cheng G, Coolen LM, Padmanabhan V, Goodman RL, Lehman MN (2010) The kisspeptin/
neurokinin B/dynorphin (KNDy) cell population of the arcuate nucleus: sex differences and
effects of prenatal testosterone in sheep. Endocrinology 151:301–311
Clark ME, Mellon PL (1995) The POU homeodomain transcription factor Oct-1 is essential for
activity of the gonadotropin-releasing hormone neuron-specific enhancer. Mol Cell Biol
15:6169–6177
Clarkson J, Herbison AE (2006) Postnatal development of kisspeptin neurons in mouse hypothala-
mus; sexual dimorphism and projections to gonadotropin-releasing hormone neurons. Endocri-
nology 147:5817–5825
Clarkson J, Boon WC, Simpson ER, Herbison AE (2009) Postnatal development of an estradiol-
kisspeptin positive feedback mechanism implicated in puberty onset. Endocrinology
150:3214–3220
Constantin S, Caraty A, Wray S, Duittoz AH (2009) Development of gonadotropin-releasing
hormone-1 secretion in mouse nasal explants. Endocrinology 150:3221–3227
d’Anglemont de Tassigny X, Colledge WH (2010) The role of kisspeptin signaling in reproduction.
Physiology 25:207–217
d’Anglemont de Tassigny X, Fagg LA, Dixon JP, Day K, Leitch HG, Hendrick AG, Zahn D,
Franceschini I, Caraty A, Carlton MB, Aparicio SA, Colledge WH (2007) Hypogonadotropic
hypogonadism in mice lacking a functional Kiss1 gene. Proc Natl Acad Sci U S A
104:10714–10719
Dawlaty MM, Ganz K, Powell BE, Hu YC, Markoulaki S, Cheng AW, Gao Q, Kim J, Choi SW,
Page DC, Jaenisch R (2011) Tet1 is dispensable for maintaining pluripotency and its loss is
compatible with embryonic and postnatal development. Cell Stem Cell 9:166–175
Dawlaty MM, Breiling A, Le T, Raddatz G, Barrasa MI, Cheng AW, Gao Q, Powell BE, Li Z,
Xu M, Faull KF, Lyko F, Jaenisch R (2013) Combined deficiency of Tet1 and Tet2 causes
epigenetic abnormalities but is compatible with postnatal development. Dev Cell 24:310–323
de Roux N, Genin E, Carel JC, Matsuda F, Chaussain JL, Milgrom E (2003) Hypogonadotropic
hypogonadism due to loss of function of the KiSS1-derived peptide receptor GPR54. Proc Natl
Acad Sci U S A 100:10972–10976
Deaton AM, Bird A (2011) CpG islands and the regulation of transcription. Genes Dev
25:1010–1022
Deplus R, Delatte B, Schwinn MK, Defrance M, Méndez J, Murphy N, Dawson MA, Volkmar M,
Putmans P, Calonne E, Shih AH, Levine RL, Bernard O, Mercher T, Solary E, Urh M, Daniels
DL, Fuks F (2013) TET2 and TET3 regulate GlcNAcylation and H3K4 methylation through
OGT and SET1/COMPASS. EMBO J 32:645–655
Desroziers E, Droguerre M, Bentsen AH, Robert V, Mikkelsen JD, Caraty A, Tillet Y, Duittoz A,
Franceschini I (2012a) Embryonic development of kisspeptin neurones in rat.
J Neuroendocrinol 24:1284–1295
Desroziers E, Mikkelsen JD, Duittoz A, Franceschini I (2012b) Kisspeptin-immunoreactivity
changes in a sex- and hypothalamic-region-specific manner across rat postnatal development.
J Neuroendocrinol 24:1154–1165
Dhillo WS, Chaudhri OB, Thompson EL, Murphy KG, Patterson M, Ramachandran R, Nijher GK,
Amber V, Kokkinos A, Donaldson M, Ghatei MA, Bloom SR (2007) Kisspeptin-54 stimulates
gonadotropin release most potently during the preovulatory phase of the menstrual cycle in
women. J Clin Endocrinol Metab 92:3958–3966
Döhler KD, Hancke JL, Srivastava SS, Hofmann C, Shryne JE, Gorski RA (1984) Participation of
estrogens in female sexual differentiation of the brain; neuroanatomical, neuroendocrine and
behavioral evidence. Prog Brain Res 61:99–117
258 J. R. Kurian and E. Terasawa
Dubois SL, Acosta-Martínez M, DeJoseph MR, Wolfe A, Radovick S, Boehm U, Urban JH, Levine
JE (2015) Positive, but not negative feedback actions of estradiol in adult female mice require
estrogen receptor α in kisspeptin neurons. Endocrinology 156:1111–1120
El Majdoubi M, Sahu A, Ramaswamy S, Plant TM (2000) Neuropeptide Y: a hypothalamic brake
restraining the onset of puberty in primates. Proc Natl Acad Sci U S A 97:6179–6184
Franceschini I, Lomet D, Cateau M, Delsol G, Tillet Y, Caraty A (2006) Kisspeptin immunoreac-
tive cells of the ovine preoptic area and arcuate nucleus co-express estrogen receptor alpha.
Neurosci Lett 401:225–230
Fueshko S, Wray S (1994) LHRH cells migrate on peripherin fibers in embryonic olfactory explant
cultures: an in vitro model for neurophilic neuronal migration. Dev Biol 166:331–348
Funes S, Hedrick JA, Vassileva G, Markowitz L, Abbondanzo S, Golovko A, Yang S, Monsma FJ,
Gustafson EL (2003) The KiSS-1 receptor GPR54 is essential for the development of the murine
reproductive system. Biochem Biophys Res Commun 312:1357–1363
Gan L, Ni PY, Ge Y, Xiao YF, Sun CY, Deng L, Zhang W, Wu SS, Liu Y, Jiang W, Xin HB (2012)
Histone deacetylases regulate gonadotropin-releasing hormone I gene expression via
modulating Otx2-driven transcriptional activity. PLoS One 7:e39770
Garcia JP, Guerriero KA, Keen KL, Kenealy BP, Seminara SB, Terasawa E (2017) Kisspeptin and
neurokinin B signaling network underlies the pubertal increase in GnRH release in female
rhesus monkeys. Endocrinology 158:3269–3280
Garcia JP, Keen KL, Kenealy BP, Seminara SB, Terasawa E (2018) Role of kisspeptin and
neurokinin B signaling in male rhesus monkey puberty. Endocrinology 159:3048–3060
Gardiner-Garden M, Frommer M (1987) CpG islands in vertebrate genomes. J Mol Biol
196:261–282
Gill JC, Navarro VM, Kwong C, Noel SD, Martin C, Xu S, Clifton DK, Carroll RS, Steiner RA,
Kaiser UB (2012a) Increased neurokinin B (Tac2) expression in the mouse arcuate nucleus is an
early marker of pubertal onset with differential sensitivity to sex steroid-negative feedback than
Kiss1. Endocrinology 153:4883–4893
Gill J, Kwong C, Clark E, Carroll RS, Shi YG, Kaiser UB (2012b) A role for the histone
demethylase LSD1 in controlling the timing of pubertal onset. In: Abstracts for Endocrine
Society Annual Meeting, Huston, TX, No. OR12–3
Givens ML, Rave-Harel N, Goonewardena VD, Kurotani R, Berdy SE, Swan CH, Rubenstein JL,
Robert B, Mellon PL (2005) Developmental regulation of gonadotropin-releasing hormone
gene expression by the MSX and DLX homeodomain protein families. J Biol Chem
280:19156–19165
Glaser S, Schaft J, Lubitz S, Vintersten K, van der Hoeven F, Tufteland KR, Aasland R,
Anastassiadis K, Ang SL, Stewart AF (2006) Multiple epigenetic maintenance factors
implicated by the loss of Mll2 in mouse development. Development 133:1423–1432
Goodman RL, Coolen LM, Lehman MN (2014) A role for neurokinin B in pulsatile GnRH
secretion in the ewe. Neuroendocrinology 99:18–32
Gorski RA (1985) Sexual dimorphisms of the brain. J Anim Sci 61(Suppl 3):38–61
Gottsch ML, Cunningham MJ, Smith JT, Popa SM, Acohido BV, Crowley WF, Seminara S, Clifton
DK, Steiner RA (2004) A role for kisspeptins in the regulation of gonadotropin secretion in the
mouse. Endocrinology 145:4073–4077
Guerriero KA, Keen KL, Millar RP, Terasawa E (2012) Developmental changes in GnRH release in
response to kisspeptin agonist and antagonist in female Rhesus monkeys (Macaca mulatta):
implication for the mechanism of puberty. Endocrinology 153:825–836
Guo JU, Su Y, Shin JH, Shin J, Li H, Xie B, Zhong C, Hu S, Le T, Fan G, Zhu H, Chang Q, Gao Y,
Ming GL, Song H (2014) Distribution, recognition and regulation of non-CpG methylation in
the adult mammalian brain. Nat Neurosci 17:215–222
Hahn MA, Qiu R, Wu X, Li AX, Zhang H, Wang J, Jui J, Jin SG, Jiang Y, Pfeifer GP, Lu Q (2013)
Dynamics of 5-hydroxymethylcytosine and chromatin marks in mammalian neurogenesis. Cell
Rep 3:291–300
Herbison AE (2016) Control of puberty onset and fertility by gonadotropin-releasing hormone
neurons. Nat Rev Endocrinol 12:452–466
9 Epigenetic Regulation of the GnRH and Kiss1 Genes: Developmental Perspectives 259
Semaan SJ, Kauffman AS (2015) Daily successive changes in reproductive gene expression and
neuronal activation in the brains of pubertal female mice. Mol Cell Endocrinol 401:84–97
Semaan SJ, Murray EK, Poling MC, Dhamija S, Forger NG, Kauffman AS (2010) BAX-dependent
and BAX-independent regulation of Kiss1 neuron development in mice. Endocrinology
151:5807–5817
Semaan SJ, Dhamija S, Kim J, Ku EC, Kauffman AS (2012) Assessment of epigenetic
contributions to sexually-dimorphic Kiss1 expression in the anteroventral periventricular
nucleus of mice. Endocrinology 153:1875–1886
Seminara SB, Messager S, Chatzidaki EE, Thresher RR, Acierno JS Jr, Shagoury JK, Bo-Abbas Y,
Kuohung W, Schwinof KM, Hendrick AG, Zahn D, Dixon J, Kaiser UB, Slaugenhaupt SA,
Gusella JF, O’Rahilly S, Carlton MB, Crowley WF Jr, Aparicio SA, Colledge WH (2003) The
GPR54 gene as a regulator of puberty. N Engl J Med 349:1614–1627
Shahab M, Mastronardi C, Seminara SB, Crowley WF, Ojeda SR, Plant TM (2005) Increased
hypothalamic GPR54 signaling: a potential mechanism for initiation of puberty in primates.
Proc Natl Acad Sci U S A 102:2129–2134
Shahab M, Lippincott M, Chan YM, Davies A, Merino PM, Plummer L, Mericq V, Seminara S
(2018) Discordance in the dependence on kisspeptin signaling in mini puberty vs adolescent
puberty: human genetic evidence. J Clin Endocrinol Metab 103:1273–1276
Shen X, Liu Y, Hsu YJ, Fujiwara Y, Kim J, Mao X, Yuan GC, Orkin SH (2008) EZH1 mediates
methylation on histone H3 lysine 27 and complements EZH2 in maintaining stem cell identity
and executing pluripotency. Mol Cell 32:491–502
Simerly RB (1998) Organization and regulation of sexually dimorphic neuroendocrine pathways.
Behav Brain Res 92:195–203
Simerly RB (2002) Wired for reproduction: organization and development of sexually dimorphic
circuits in the mammalian forebrain. Annu Rev Neurosci 25:507–536
Smith JT (2008) Kisspeptin signalling in the brain: steroid regulation in the rodent and ewe. Brain
Res Rev 57:288–298
Smith JT (2013) Sex steroid regulation of kisspeptin circuits. Adv Exp Med Biol 784:275–295
Smith JT, Cunningham MJ, Rissman EF, Clifton DK, Steiner RA (2005) Regulation of Kiss1 gene
expression in the brain of the female mouse. Endocrinology 146:3686–3692
Smith JT, Li Q, Pereira A, Clarke IJ (2009) Kisspeptin neurons in the ovine arcuate nucleus and
preoptic area are involved in the preovulatory luteinizing hormone surge. Endocrinology
150:5530–5538
Smith JT, Shahab M, Pereira A, Pau KY, Clarke IJ (2010) Hypothalamic expression of KISS1 and
gonadotropin inhibitory hormone genes during the menstrual cycle of a non-human primate.
Biol Reprod 83:568–577
Stresemann C, Lyko F (2008) Modes of action of the DNA methyltransferase inhibitors azacytidine
and decitabine. Int J Cancer 123:8–13
Suberbielle E, Sanchez PE, Kravitz AV, Wang X, Ho K, Eilertson K, Devidze N, Kreitzer AC,
Mucke L (2013) Physiologic brain activity causes DNA double-strand breaks in neurons, with
exacerbation by amyloid-β. Nat Neurosci 16:613–621
Svedruzic ZM (2008) Mammalian cytosine DNA methyltransferase Dnmt1: enzymatic mechanism,
novel mechanism-based inhibitors, and RNA-directed DNA methylation. Curr Med Chem
15:92–106
Terasawa E (2001) Luteinizing hormone-releasing hormone (LHRH) neurons: mechanism of
pulsatile LHRH release. Vitam Horm 63:91–129
Terasawa E, Quanbeck CD, Schulz CA, Burich AJ, Luchansky LL, Claude P (1993) A primary cell
culture system of luteinizing hormone releasing hormone (LHRH) neurons derived from fetal
olfactory placode in the rhesus monkey. Endocrinology 133:2379–2390
Terasawa E, Keen KL, Mogi K, Claude P (1999) Pulsatile release of luteinizing hormone-releasing
hormone (LHRH) in cultured LHRH neurons derived from the embryonic olfactory placode of
the rhesus monkey. Endocrinology 140:1432–1441
Terasawa E, Busser BW, Luchansky LL, Sherwood NM, Jennes L, Millar RP, Glucksman MJ,
Roberts JL (2001) Presence of luteinizing hormone-releasing hormone fragments in the rhesus
monkey forebrain. J Comp Neurol 439:491–504
264 J. R. Kurian and E. Terasawa
Terasawa E, Garcia JP, Seminara SB, Keen KL (2018) Role of kisspeptin and neurokinin B in
puberty in female non-humanprimates. Front Endocrinol (Lausanne) 9:148
Tomikawa J, Uenoyama Y, Ozawa M, Fukanuma T, Takase K, Goto T, Abe H, Ieda N, Minabe S,
Deura C, Inoue N, Sanbo M, Tomita K, Hirabayashi M, Tanaka S, Imamura T, Okamura H,
Maeda K, Tsukamura H (2012) Epigenetic regulation of Kiss1 gene expression mediating
estrogen-positive feedback action in the mouse brain. Proc Natl Acad Sci U S A 109:1294–1301
Topaloglu AK, Tello JA, Kotan LD, Ozbek MN, Yilmaz MB, Erdogan S, Gurbuz F, Temiz F,
Millar RP, Yuksel B (2012) Inactivating KISS1 mutation and hypogonadotropic hypogonadism.
N Engl J Med 366:629–635
Toro CA, Wright H, Aylwin CF, Ojeda SR, Lomniczi A (2018) Trithorax dependent changes in
chromatin landscape at enhancer and promoter regions drive female puberty. Nat Commun 9:57
Uenoyama Y, Tomikawa J, Inoue N, Goto T, Minabe S, Ieda N, Nakamura S, Watanabe Y,
Ikegami K, Matsuda F, Ohkura S, Maeda K, Tsukamura H (2016) Molecular and epigenetic
mechanism regulating hypothalamic Kiss1 gene expression in mammals. Neuroendocrinology
103:640–649
Vargas Trujillo M, Kalil B, Ramaswamy S, Plant TM (2017) Estradiol upregulates kisspeptin
expression in the preoptic area of both the male and female Rhesus monkey (Macaca mulatta):
implications for the hypothalamic control of ovulation in highly evolved primates. Neuroendo-
crinology 105:77–89
Wallen K (2005) Hormonal influences on sexually differentiated behavior in nonhuman primates.
Front Neuroendocrinol 26:7–26
Wallen K (2009) The organizational hypothesis: reflections on the 50th anniversary of the publica-
tion of Phoenix, Goy, Gerall, and Young (1959). Horm Behav 55:561–565
Watanabe Y, Uenoyama Y, Suzuki J, Takase K, Suetomi Y, Ohkura S, Inoue N, Maeda KI,
Tsukamura H (2014) Oestrogen-induced activation of preoptic kisspeptin neurones may be
involved in the luteinising hormone surge in male and female Japanese monkeys.
J Neuroendocrinol 26:909–917
Whyte DB, Lawson MA, Belsham DD, Eraly SA, Bond CT, Adelman JP, Mellon PL (1995) A
neuron-specific enhancer targets expression of the gonadotropin-releasing hormone gene to
hypothalamic neurosecretory neurons. Mol Endocrinol 9:467–477
Wu SC, Zhang Y (2010) Active DNA demethylation: many roads lead to Rome. Nat Rev Mol Cell
Biol 11:607–620
Xie H, Hoffmann HM, Meadows JD, Mayo SL, Trang C, Leming SS, Maruggi C, Davis SW,
Larder R, Mellon PL (2015) Homeodomain proteins six3 and six6 regulate gonadotrope-specific
genes during pituitary development. Mol Endocrinol 29:842–855
Yagi H, Deguchi K, Aono A, Tani Y, Kishimoto T, Komori T (1998) Growth disturbance in fetal
liver hematopoiesis of Mll-mutant mice. Blood 92:108–117
Yamaguchi S, Hong K, Liu R, Shen L, Inoue A, Diep D, Zhang K, Zhang Y (2012) Tet1 controls
meiosis by regulating meiotic gene expression. Nature 492:443–447
Yang H, Lin H, Xu H, Zhang L, Cheng L, Wen B, Shou J, Guan K, Xiong Y, Ye D (2014)
TET-catalyzed 5-methylcytosine hydroxylation is dynamically regulated by metabolites. Cell
Res 24:1017–1020
Yeo SH, Clarkson J, Herbison AE (2014) Kisspeptin-gpr54 signaling at the GnRH neuron is
necessary for negative feedback regulation of luteinizing hormone secretion in female mice.
Neuroendocrinology 100:191–197
Yu BD, Hess JL, Horning SE, Brown GA, Korsmeyer SJ (1995) Altered Hox expression and
segmental identity in Mll-mutant mice. Nature 378:505–508
Imprinted Genes and Hypothalamic
Function 10
Michela Pulix and Antonius Plagge
Abstract
Genomic imprinting, a specific type of inherited epigenetic regulation, controls a
number of genes with important functions in the hypothalamus and its associated
neuroendocrinology. Imprinted genes are expressed monoallelically, dependent
on the parental origin of the allele, whereby maternally or paternally inherited
DNA methylation marks determine gene activity. In this chapter, we provide an
overview of imprinted genes with roles in the hypothalamic regulation of whole-
body energy homeostasis, neuroendocrine hormone functions, and circadian
rhythmicity. As an example for imprinted genes that impact on the central
regulation of energy balance, we describe the Gnas locus in more detail, since
it generates parental allele-specific gene products with antagonistic roles in
energy expenditure and sympathetic nervous system activity. Several human
neuroendocrine disorders are caused by defects in imprinted genes as described
here for Prader–Willi Syndrome and Angelman Syndrome. Furthermore, we
outline methodological approaches for the investigation of DNA methylation
marks and allelic gene expression, including bisulfite conversion of DNA and
pyrosequencing. Recent technological advances to resolve methylation and alle-
lic expression at the single-cell level are introduced.
M. Pulix
Cellular and Molecular Physiology, Institute of Translational Medicine, University of Liverpool,
Liverpool, UK
Department of Neuroscience, Sheffield Institute for Translational Neuroscience, University of
Sheffield, Sheffield, UK
A. Plagge (*)
Cellular and Molecular Physiology, Institute of Translational Medicine, University of Liverpool,
Liverpool, UK
e-mail: [email protected]
Keywords
Genomic imprinting · Hypothalamus · Gnas · Energy homeostasis · Prader–Willi
syndrome · Angelman syndrome · DNA methylation · Bisulfite conversion of
DNA · Pyrosequencing
10.1 Introduction
The term genomic imprinting refers to the expression of certain genes from only one
of the two parental alleles. This is dependent on the parental origin, i.e., some genes
are monoallelically expressed from the maternal, others from the paternal allele,
while the opposite alleles are silenced. The allelic silencing is already determined in
the parental germ cells through epigenetic “marks” and maintained in the somatic
tissues of the offspring (Fig. 10.1). To maintain this parent-of-origin-specific gene
expression, the epigenetic “marks” are erased in the primordial germline cells of the
developing offspring and reset according to the sex of the offspring. This process
ensures that the maternal or paternal allelic expression, respectively, of an imprinted
gene is inherited from generation to generation. To date, circa 150 imprinted genes
have been confirmed in humans and mice, but imprinted gene expression also exists
in other mammals (http://www.mousebook.org/imprinting-gene-list; http://igc.
otago.ac.nz/home.html). Most imprinted genes occur in conserved clusters, but the
imprinting of some genes is species specific.
Many imprinted genes have essential roles in embryonic development and pla-
cental functions (Plasschaert and Bartolomei 2014). However, they also play impor-
tant roles in the brain, including regulation of neurogenesis, synaptic transmission,
and behavior (Perez et al. 2015; Peters 2014). Within the brain, the hypothalamus
has proven to be the main site of expression of imprinted genes, and from studies of
knockout mouse models as well as human inherited disorders, it is evident that they
influence neuroendocrine pathways, energy homeostasis, circadian rhythm, and
social interactions (Ivanova and Kelsey 2011; Perez et al. 2015). In this chapter,
we provide an overview of the hypothalamic functions of imprinted genes and
highlight a few disease-relevant examples. In addition, we discuss techniques used
in the identification and epigenetic analysis of imprinted genes and indicate currently
open questions and trends in the field.
Erasure of DNA
methylation
imprinting marks in
Oocyte primordial germ cells
of embryo
Sperm Embryo
Maintenance of
germline-derived DNA
methylation imprinting
marks during genome-
wide demethylation
events in early embryo
Zygote
Blastocyst
Fig. 10.1 Inheritance cycle of DNA methylation imprinting marks through the germline and
maintenance in somatic tissues. The genomic DNA methylation imprints are established in the
germline cells in a sex-specific manner and transmitted to the zygote, where they resist
reprogramming events and global changes in DNA methylation, which take place after fertilization.
While the genomic imprinting marks are maintained in somatic cells throughout life, they are erased
in the primordial germ cells of the developing embryo and re-established de novo according to the
sex of the organism, to be transmitted to the next generation. Re-establishment of imprinting marks
in gametogenesis occurs prenatally in males and at postnatal stages in females [Adapted from
Bartolomei and Ferguson-Smith (2011)]
such is not sufficient for normal mammalian development, that the two parental
genomes make different functional contributions to embryonic development and that
there must be some kind of “mark,” which mediates the distinct functions of the
maternal and paternal genomes, respectively. Thus, genomic imprinting became a
paradigm that drove research into “epigenetic” mechanisms (Ferguson-Smith 2011).
We know today that those epigenetic “marks” represent methylation of genomic
DNA, histone modifications, and noncoding RNA expression.
In the majority of cases, imprinted genes are typically found in clusters of 3–12
genes spread over 20 Kb–3.7 Mb of DNA (Ferguson-Smith 2011; Peters, 2014) and
268 M. Pulix and A. Plagge
they are usually regulated by the presence of DNA methylation on one parental allele
in so-called Differentially Methylated Regions (DMRs). These regions are
characterized by a high density of CpG dinucleotides (above 55%), termed CpG
islands (CGIs). Here, methylation of cytosines occurs at the C5 position of the
pyrimidine ring and is catalyzed by DNA methyltransferases (DNMTs) (Lyko
2018). A DMR is termed an Imprinting Control Region (ICR), if the differential
methylation is established already in the germ cells and if functional studies have
proven that it controls the imprinting of associated genes within a cluster. Thus, ICRs
regulate gene expression and chromatin structure over long distances (Ferguson-
Smith 2011). Maternally methylated DMRs and ICRs are usually located at
promoters while paternally methylated DMRs and ICRs are located in intergenic
regions (Bartolomei and Ferguson-Smith 2011).
DNA methylation is established during gametogenesis, resists the extensive
genomic reprogramming that occurs after fertilization and is maintained in the
offspring throughout life (Fig. 10.1). For inheritance of imprinted gene expression
to the next generation, the primordial germ cells of an embryo undergo a phase of
erasure of DNA methylation, followed by re-establishment of methylation at DMRs/
ICRs according to the sex of the individual and mediated by DNMT3A and the
accessory protein DNMT3L (Bartolomei and Ferguson-Smith 2011; Plasschaert and
Bartolomei 2014). How the ICRs are recognized within the genome is still unknown,
but it is clear that several factors are involved in the maintenance of DNA methyla-
tion, including ZINC FINGER PROTEIN HOMOLOG 57 (ZFP57) and Develop-
mental pluripotency-associated 3 (Dppa3). Mutations in these genes result in loss of
imprinting at several loci during early embryogenesis, which in turn causes a range
of disease symptoms (Ferguson-Smith 2011; Plasschaert and Bartolomei 2014).
Before discussing techniques and methods used in the analysis of allelic expres-
sion and epigenetic regulation of imprinted genes, we first provide an overview
about the functional roles of imprinted genes in the hypothalamus and, where
applicable, their involvement in human neurodevelopmental disorders. Then, in
the context of the Prader–Willi syndrome (PWS)/Angelman syndrome (AS) imprint-
ing cluster and the Gnas locus, we discuss the structure and organization of regu-
latory elements of imprinted genes.
When the first imprinted genes (Igf2, Igf2r, and H19) were identified, they were
found to affect mainly embryonic growth, placental function, and nutrient supply
and demand at the maternal–fetal interface (Ferguson-Smith 2011; Peters 2014;
Plasschaert and Bartolomei 2014). However, it soon became apparent that a number
of imprinted genes were expressed in defined brain regions, and many of them
showed high expression levels specifically in the hypothalamus (Ivanova and Kelsey
2011). Phenotype analyses of knockout mouse models and clinical case studies of
human imprinting disorders revealed that many imprinted genes continue to impact
growth and whole-body energy homeostasis beyond embryonic development and
10 Imprinted Genes and Hypothalamic Function 269
into adulthood via influencing neuroendocrine pathways and the autonomic nervous
system. Other core functions of the hypothalamus, including circadian rhythms,
reproductive functions, and maternal-infant care behavior are also deregulated
upon the deficiency of some imprinted genes. We will first describe the hypotha-
lamic roles of the Gnas and the PWS-AS loci since they have been analyzed in some
detail, before summarizing findings related to other imprinted genes.
A) Nesp55
Gαs
MMMMMMMMM MMMM
Maternal
allele:
Paternal
allele:
MMMMM
( Gαs
)
XLαs
B)
MM MM MMMMM
Maternal
allele:
Fig. 10.2 Simplified schemes of the Gnas and the PWS-AS (Snrpn-Ube3a) imprinted loci. The
features of the maternal and paternal alleles are indicated in red and blue, respectively. Genes or
transcripts are named in the central part. Arrows mark transcription start sites and undulating lines
noncoding RNAs. Open and filled boxes represent noncoding and coding exons, respectively, while
black boxes indicate silenced genes. Differentially methylated regions of DNA (DMRs) are marked
by MMM. (a) For the Gnas locus, the alternatively spliced coding transcripts and proteins are
named above and below the alleles. DMRs at Nespas and Exon 1A (EXON A/B in human) are
established in the maternal germline, while methylation at Nesp occurs during early embryonic
development. Gnas is expressed biallelically, but is silenced on the paternal allele in some tissues
(hatched box). Nesp represents a coding transcript (ORF limited to Nesp exon 2). Nespas is
expressed from the unmethylated imprinting control region (ICR) of the locus. (b) The PWS-AS-
associated Snrpn-Ube3a imprinting cluster contains several genes (represented by single-exon
boxes) that show monoallelic expression in the brain. Ube3a is biallelically expressed in most
tissues, but silenced on the paternal allele in neurons (hatched box). The Snrpn long noncoding
RNA occurs in multiple, variably processed forms, including brain-specific variants that overlap
with Ube3a (interrupted undulating lines). Snord115 and Snord116 represent clusters of C/D box
snoRNAs, which are generated from the Snrpn transcript. The bipartite ICR of the human locus is
indicated as AS-IC and PWS-IC, the latter being conserved in mouse around the main Snrpn start
site. NPAP1 is not present in the murine locus, but Peg12 is imprinted in mice. Somatic DMRs at
Ndn and Mkrn3 are established during development
10 Imprinted Genes and Hypothalamic Function 271
Magel2 MAGE family ubiquitin ligase Paternal Postnatally: 10–50% mortality; Prader–Willi syndrome
regulator (endosomal protein diminished oxytocin-induced suckling
sorting and recycling) initiation after birth; reduced neonatal
hypothalamic levels of oxytocin,
arginine-vasopressin, and orexin-A
Adult stage: increased fat mass despite
reduced food intake and orexin-A levels,
reduced lean mass; hypothalamic leptin
resistance; reduced POMC neurons and
projections; rhythmic circadian
expression in SCN and altered/reduced
circadian activity patterns; and abnormal
behavior
Ndn MAGE family ubiquitin ligase Paternal High rate of early neonatal mortality due Prader–Willi syndrome
(Necdin) regulator to respiratory deficiencies (brainstem
Imprinted Genes and Hypothalamic Function
nuclei in the medulla, e.g., the raphe pallidus, nucleus of the solitary tract, as well as
the intermediolateral layer of the spinal cord (Krechowec et al. 2012; Plagge et al.
2004). Expression of Gnasxl in peripheral tissues, for example, muscle tissues,
becomes silenced postnatally with only a few exemptions like the adrenal medulla
(Krechowec et al. 2012; Xie et al. 2006). Newborn XLαs-deficient pups become
growth retarded, take in less milk and display hypoglycemia and lack of lipid
reserves in adipose tissue, all of which results in a high mortality rate in the first
week (Table 10.1; Plagge et al. 2004; Xie et al. 2006). Those Gnasxl knockout mice
that survive to weaning age, develop into fertile adults, but remain underweight and
lean throughout life, despite relatively increased food intake. Furthermore, their
phenotype consists of increased energy expenditure, metabolic rate, body tempera-
ture, blood pressure, and heart rate, which is associated with elevated SNS outflow
(Nunn et al. 2013; Xie et al. 2006). They also show improved glucose tolerance and
insulin sensitivity (Xie et al. 2006). The precise neural pathways through which
Gnasxl regulates SNS outflow and energy homeostasis remain unresolved, as are the
neural receptors that signal via XLαs.
Human disease symptoms related to loss of GNASXL function have been difficult
to ascertain since mutations that specifically disable the first GNASXL exon have not
been identified. However, paternal transmission of mutations in shared downstream
exons 2–13 and maternal uniparental disomies that encompass the GNAS locus on
chromosome 20, which result in two maternal alleles and lack of a paternally inherited
allele, indicate pre-/neonatal symptoms (Table 10.1). These include intrauterine
growth retardation, low birth weight, postnatal growth retardation, and feeding
problems (Genevieve et al. 2005; Kawashima et al. 2018; Mulchandani et al. 2016;
Richard et al. 2013). Although no adult metabolic data of such patients have been
reported, the neonatal symptoms are reminiscent of the Gnasxl knockout mouse
phenotype (Plagge et al. 2004). Taken together, these findings show that XLαs exerts
an important role in neonatal physiology in addition to its regulation of adult energy
homeostasis, which to date has only been confirmed in mice.
In adults, XLαs acts in the brain to inhibit SNS outflow and energy expenditure,
while its variant Gαs functions to stimulate SNS activity and metabolic rate. These
antagonistic roles are correlated on the epigenetic level with genomic imprinting and
monoallelic transcription from opposite parental alleles. Gnasxl is expressed exclu-
sively from the paternal allele, while Gnas is only imprinted in specific hypothalamic
neurons, where it is transcribed from the maternal allele. Why epigenetic regulation
through genomic imprinting of crucial growth and energy homeostasis functions
evolved in mammals is beyond the scope of this chapter, but has been extensively
discussed in the literature (Haig 2004; Peters 2014).
(Bischof et al. 2007; Kozlov et al. 2007). It is noteworthy that the observation of
reduced food intake is contrary to PWS symptoms. Hypothalamic POMC neurons
and their projections were found to be deficient in knockout mice (Maillard et al.
2016; Mercer et al. 2013), and leptin resistance might be related to reduced leptin
receptor recycling from endosomes to the plasma membrane (Wijesuriya et al.
2017). The Magel2 protein is also robustly expressed in the SCN and its levels
fluctuate with circadian rhythmicity (Kozlov et al. 2007). Consistent with SCN
function, the knockout mice show overall reduced wheel-running activity and
disturbed activity patterns under constant darkness (Kozlov et al. 2007; Tacer and
Potts 2017).
Similar to Magel2, Necdin (Ndn) constitutes another member of the Mage family
of proteins (Tacer and Potts 2017). Necdin is expressed in a number of brain regions,
and it has roles in neuronal survival, migration, differentiation, and neurite formation
(Table 10.1). Ndn knockout mice have neonatal respiratory problems, which lead to
a substantial rate of mortality shortly after birth (Gerard et al. 1999; Muscatelli et al.
2000). Respiratory disturbances are also observed in PWS and they are due to
defects in the serotoninergic system of the brainstem inspiratory rhythm generator
(Matarazzo et al. 2017). In the hypothalamus, lack of Ndn causes reduced numbers
of oxytocinergic neurons in the PVN and gonadotropin-releasing hormone neurons
in the preoptic area (Miller et al. 2009; Muscatelli et al. 2000). Deficits in the latter
neuron population might be responsible for the hypogonadism observed in PWS.
Necdin knockout mice present symptoms of hypothyroidism due to reduced levels of
thyrotropin-releasing hormone in the PVN (Hasegawa et al. 2012). Furthermore,
Necdin has a role in axon outgrowth as indicated by aberrant axon bundles in the
anterior hypothalamus and other brain regions (Lee et al. 2005). On the molecular
level, Necdin appears to interact with a diverse set of binding partners, including
cytoskeleton-associated proteins as well as transcription factors (Hasegawa et al.
2012; Lee et al. 2005).
The box C/D small nucleolar RNA cluster Snord116 also has a crucial role in
PWS (Cavaille 2017). Patients with microdeletions of this cluster, but intact
NECDIN and MAGEL2 genes, present with PWS symptoms, and knockout mouse
models also show some of these disease characteristics (Bieth et al. 2015; Ding et al.
2008; Skryabin et al. 2007). Postnatal growth retardation is a shared feature in
humans and mice with paternally inherited Snord116 deletions. However, knockout
mice remain lean and underweight throughout life due to increased energy expendi-
ture, while SNORD116-mutant patients develop the typical PWS hyperphagic obe-
sity toward adulthood (Bieth et al. 2015; Ding et al. 2008; Skryabin et al. 2007).
Interestingly, when the Snord116 cluster is deleted in the mediobasal hypothalamus
of adult mice via an adenovirus Cre-recombinase vector injection, the mutant mice
then develop hyperphagia and obesity reminiscent of PWS patients (Polex-Wolf
et al. 2018). These findings highlight physiological differences between humans and
mice and potentially indicate species-specific compensatory abilities to inherited
gene mutations during development.
10 Imprinted Genes and Hypothalamic Function 279
Although less well characterized, four other imprinted genes have roles in the
hypothalamus, which are briefly summarized here. A cluster of imprinted genes on
mouse chromosome 12 (human chromosome 14) contains two protein-coding genes
with hypothalamic functions, Dlk1 (Protein delta homolog 1) and Dio3 (Thyroxine
5-deiodinase), both expressed from the paternal allele (Table 10.1). The Dlk1
protein has structural similarity to Notch ligands, but its precise signaling mecha-
nism is unknown. Dlk1 is prominently expressed in several hypothalamic nuclei
(arcuate nucleus, PVN, DMH, and SCN), and it has been co-localized in oxytocin,
arginine–vasopressin, orexin, NPY, and dynorphin-positive neurons (Meister et al.
2013; Villanueva et al. 2012). Dlk1 involvement in embryonic development and
adiposity has been well characterized, but its hypothalamus-specific functions are
poorly understood. A mutation of DLK1 in humans leads to central precocious
puberty, possibly via GnRH-neuron regulation, and increased fat mass (Dauber
et al. 2017). Dio3 shows monoallelic expression from the paternal allele in the
neonatal hypothalamus (Martinez et al. 2014). Knockout mice display postnatal
growth retardation, which lasts into adulthood, deregulated thyroid hormone levels
and disturbances in the hypothalamic–pituitary–thyroid axis (Table 10.1). Reduced
adiposity is associated with changes in hypothalamic gene expression of the leptin–
melanocortin pathway, increased energy expenditure, and locomotion (Wu et al.
2017). Activity patterns of knockout mice also showed circadian changes, which
might be associated with Dio3 expression in the SCN. Additional behavioral
abnormalities relating to aggression and maternal care of the knockout mice are
associated with changes in the hypothalamic oxytocin and arginine–vasopressin
systems (Stohn et al. 2018). The third gene is the voltage-gated potassium channel
gene Kcnq1, which is located in a large imprinting cluster on mouse chromosome
7 (human chromosome 11) and is expressed from the maternal allele. It has a crucial
role in the hypothalamic–pituitary growth hormone axis. The potassium channel is
expressed in hypothalamic growth hormone-releasing hormone (GHRH) neurons
and pituitary somatotropes. KCNQ1 mutations in humans cause growth hormone
deficiency and gingival fibromatosis among other symptoms (Table 10.1)
(Tommiska et al. 2017). The fourth imprinted gene associated with hypothalamic
functions is the Paternally expressed gene 3 (Peg3), a Zinc finger-containing
DNA-binding factor that is widely expressed in the hypothalamus (Li et al. 1999).
Peg3 knockout mice have reduced body weight at all stages of life, but are
hypometabolic, show a proportionately increased fat mass, lower body temperature,
leptin resistance and changes in hypothalamic NPY, POMC, MCH, and Orexin
neuropeptide expression (Curley et al. 2005). Whether knockout females show
deficiencies in maternal behavior toward their offspring (e.g., nest building, pup
retrieval, nursing) that are linked to a reduced number of hypothalamic oxytocin
neurons, is controversial (Li et al. 1999; Denizot et al. 2016).
280 M. Pulix and A. Plagge
NH2
O N O
H
Cytosine HN
NH2 O N
H
CH3
N Uracil
O N
H
5-Methylcytosine
5’-ACCTAAGGTCGTTACCCTACGTTAGGGCGTAGCAACGTGGACGCTATCTAGCTGCATCGAT-3’
5’-ATTTAAGGTTGTTATTTTACGTTAGGGTGTAGTAACGTGGATGTTATTTAGTTGTATTGAT-3’
Fig. 10.3 Bisulfite conversion of genomic DNA. The treatment with NaHSO3 induces the deami-
nation of cytosine into uracil residues. This conversion will not happen, if the cytosine is methylated
or hydroxy-methylated in the carbon 5 position. Following bisulfite treatment, the sense and
antisense strands of the original double-stranded DNA are no longer complementary. PCR amplifi-
cation of a strand will now lead to an exchange of uracil residues to thymine and the DNA sequence
will then be characterized by an abundance of A and T. Only the methylated sites will retain
cytosine residues. The primers for bisulfite sequencing need to be complementary to the converted
sequence (orange and green arrows). Black lollipops indicate methylated CpG sites
originally methylated CpGs as they are conserved after bisulfite treatment and PCR
amplification and then cut by the restriction endonuclease. By contrast, bisulfite
conversion of an unmethylated CpG leads to loss of the restriction site. If a CpG is
fully methylated, the PCR product will be completely digested, and vice versa
(Fig. 10.4b). Quantification of partially methylated sites will require blotting of the
gel, hybridization, and signal detection with a probe.
The presence of both cytosine and thymine in the same position following
bisulfite treatment and PCR amplification can indicate partial methylation or poten-
tially incomplete bisulfite conversion. In the bisulfite sequencing approach, incom-
plete conversion can be recognized through the presence of remaining C residues
outside of CpG dinucleotides—all of these should have been converted. A potential
282 M. Pulix and A. Plagge
differentially
methylated
Predominantly
unmethylated
BstUI digest
1000 1000
Kidney
30
20
10
-10
E S C A A T A T C G A T C C A G A C T A C A T A G C A T A T C G A C A A T C A C A C T C T A T A T A C G A T C
5 10 15 20 25 30 35 40 45 50
Fig. 10.4 Analysis methods of bisulfite-treated and PCR-amplified genomic DNA. (a) Cloning
and Sanger sequencing of bisulfite PCR products. Scheme showing three examples of possible
outcomes for methylation patterns in an amplicon covering a CGI. Each row of dots represents one
cloned PCR product and each dot represents an individual CpG site (black ¼ methylated,
white ¼ unmethylated). The differentially methylated example, which contains methylated and
unmethylated sequences, would be typical of an imprinted DMR. Further SNP analysis would be
required to associate the methylation status with the parental origin of an allele. (b) Combined
bisulfite restriction analysis (COBRA) on untreated and bisulfite-treated genomic DNA from
hippocampus, cerebral cortex, and kidney. Restriction enzyme BstUI recognizes the site CGCG.
When a cytosine is unmethylated in the genomic DNA, it is converted to thymine following bisulfite
treatment and PCR amplification, disrupting the integrity of the restriction sites. In contrast,
10 Imprinted Genes and Hypothalamic Function 283
Fig. 10.4 (continued) methylated sequences maintain the cytosine and the restriction enzyme will
cut the PCR amplicon. This example CpG island shows the presence of both methylated and
unmethylated alleles (unpublished data). (c) Bisulfite pyrosequencing analysis of the same CGI as
in (b). A pyrogram sequencing trace of a bisulfite PCR amplicon is shown. The positions
highlighted in light blue represent CpG sites and the degree of methylation at each site is quantified
and displayed as a percentage above the highlighted peaks (unpublished data)
284 M. Pulix and A. Plagge
BRAIN KIDNEY
Fig. 10.5 SNP pyrosequencing to quantify allele-specific expression levels of a gene. Brain and
kidney cDNAs were prepared from reciprocal crosses of C57BL/6J (B6) and Japanese fancy mouse
1 (Mus musculus molossinus, JF1), and a SNP-containing amplicon analyzed via pyrosequencing.
Quantification of SNP prevalence indicates ~70% preferential maternal allele-specific expression of
Trappc9 in the brain, but equal biallelic expression in kidney. The reciprocal mouse crosses are
shown as JF1/B6 and B6/JF1, whereby the first position conventionally indicates the maternal
genotype. The SNP position is shaded in light blue
ratio of expression of the SNPs within the gene of interest determined via
pyrosequencing. In the example shown in Fig. 10.5, expression of the Trappc9
gene was found to be predominantly from the maternal allele (~70%) in brain tissue,
while in kidney both SNPs show equal biallelic expression levels.
Recently, parent-of-origin biases of gene expression in the hybrid mouse brain
have been characterized using RNAseq whole-transcriptome methods. In some
cases, specific brain subregions, including the hypothalamus and arcuate nucleus
were analyzed (Babak et al. 2015; Bonthuis et al. 2015; Huang et al. 2017; Perez
et al. 2015). These transcriptome datasets, which include previously unknown
transcripts, can be a good starting point for the investigation of imprinted genes in
the brain. While several genes, including Gnas, Dlk1, and Trappc9, show tissue- or
cell type-specific imprinting (Chen et al. 2009; Ferron et al. 2011; Perez et al. 2015),
it is often unclear whether every cell within a population exhibits the parent-of-
origin-specific, monoallelic expression. In particular, for genes with a bias toward
parental allele-specific expression (e.g., 70% maternal, 30% paternal transcripts,
based on whole tissue lysates) it remains unexplained whether such a ratio reflects
unequal biallelic expression within each cell, or whether it is due to a mixture of cells
with absolute monoallelic expression and cells with equal biallelic expression. This
question is now being addressed with single-cell technologies. There are two main
approaches to the single-cell analysis of allelic gene expression. On the one hand,
probe hybridization onto fixed cells or tissue sections is combined with microscopy
to distinguish between mono- or biallelic expression in situ (Bonthuis et al. 2015;
10 Imprinted Genes and Hypothalamic Function 285
Ginart et al. 2016; Huang et al. 2017). On the other hand, molecular genetics or
genome-wide next-generation sequencing methods are applied on isolated single cell
extracts (Angermueller et al. 2016; Cheow et al. 2015; Clark et al. 2018; Deng et al.
2014; Kelsey et al. 2017). As some of these techniques have been applied to
hypothalamic tissues, we briefly describe them here as promising approaches for
future research.
One of the methods designed to investigate allele-specific expression at the tissue
level is “single nucleotide polymorphism RNA fluorescent in situ hybridization”
(SNP-FISH) (Ginart et al. 2016; Levesque et al. 2013). The method is applied on
tissues or cells from hybrid mice (e.g., C57BL/6J Cast/EiJ), which harbor several
SNPs in the mRNA of the gene of interest. Three different pools of fluorescently
labeled oligonucleotides are then hybridized simultaneously to fixed samples: (a) a
pool of “guide” probes, which bind to the mRNA of the gene of interest outside of
the SNP regions for general detection of the relevant transcript; (b) two pools of
SNP-specific oligonucleotides with different fluorescent labels, which vary only in a
single nucleotide position to detect the parental origin of the transcript (Fig. 10.6).
These SNP-specific oligos are initially blocked by shorter “mask” oligonucleotides,
leaving the SNP-containing region free to hybridize to the perfectly matching
mRNA. Eventually, the SNP-specific oligos shed their “mask” oligos to fully anneal
to the mRNA of interest. By analyzing co-localization of fluorescent signals from the
“guide” oligos and the SNP-specific oligos, respectively, Ginart and colleagues
(Ginart et al. 2016) were able to demonstrate that, within one tissue, cell populations
with monoallelic maternal expression of the imprinted gene H19 coexisted with
neighboring cell populations that showed biallelic expression.
Another technique used to visualize the imprinted expression of genes in tissues
is RNAscope in situ hybridization. Bonthuis and colleagues (Bonthuis et al. 2015)
adapted the standard RNAscope (Advanced Cell Diagnostics, ACD) method to
investigate actively transcribed alleles of specific genes in individual cell nuclei of
fixed sections from the hypothalamic arcuate nucleus and dorsal raphe nucleus
(Bonthuis et al. 2015). Using intronic oligonucleotide probes, which detect nascent
transcripts in the nucleus before splicing occurs, they were able to determine through
imaging for the presence of one or two hybridization signals (“color substrate dots”)
within a nucleus, whether one or both alleles of a gene were expressed. In contrast to
the SNP-FISH technique, however, the RNAscope method does not provide infor-
mation on the parental origin of monoallelic expression. Also, when used on tissue
sections, the RNAscope approach has an inherent false detection rate problem, since
the plane of sectioning might result in spatial separation of the two alleles in
sequential sections. This would result in single hybridization signals within nuclei
although biallelic expression occurs. This false detection rate needs to be accounted
for by using appropriate control probes for biallelically expressed genes. The authors
estimate that this method has an error rate of ~25% nuclei showing potentially false
monoallelic expression (Bonthuis et al. 2015; Huang et al. 2017). Nevertheless, due
to the high sensitivity of the hybridization probes and the straight forward protocol,
this approach can provide a useful first assessment of the localization of cells with
mono- and biallelic expression within tissues. Especially since it can be combined
286
initial annealing
allele B RNA 5’-GCGATCTTGTCGAGCTGTCGATGTCGACT
3’-UGUGAGCUUGAUGCUAGCAUCGACCUUGACGAUCGAUGGCUAAGAUCGGAUCGUA……GAUCGCCAAGCUCGCUAGAACAGCUCGACAGCUACAGCUGACACAAGCUCUCUUUG-5’
guide oligonucleotide
allele A RNA 5’-TAGCTAGAACTGCTAGCTACCGATTCTA 5’-GCGATCTTGTCGAGCTGTCGATGTCGACT
3’-UGUGAGCUCGAUGCUAGCAUCGAUCUUGACGAUCGAUGGCUAAGAUCGGAUCGUA……GAUCGCCAAGCUCGCUAGAACAGCUCGACAGCUACAGCUGACACAAGCUCUCUUUG-5’
guide oligonucleotide
allele B RNA 5’-TAGCTGGAACTGCTAGCTACCGATTCTA 5’-GCGATCTTGTCGAGCTGTCGATGTCGACT
fully annealed
3’-UGUGAGCUCGAUGCUAGCAUCGACCUUGACGAUCGAUGGCUAAGAUCGGAUCGUA……GAUCGCCAAGCUCGCUAGAACAGCUCGACAGCUACAGCUGACACAAGCUCUCUUUG-5’
allele A
Cytoplasm
transcription sites
non-specific
Nucleus
Fig. 10.6 Schematic illustration of the SNP-FISH technique for the detection of allele-specific RNA expression in cells and tissue sections. The allelic RNA
variants of a gene of interest are shown with a SNP highlighted in red. A pool of fluorescently labeled “guide” oligonucleotides, which hybridize to non-variant
sequences of the RNA, serve as a general marker for the presence of the transcript. Allele-specific oligonucleotides covering SNP positions (SNV probes) are
labeled with a second and third fluorescent tag, respectively. Initially, only a short stretch of the SNV probe sequence (including the actual SNP position) is free
to hybridize with the target RNA, as the remainder of the oligo is bound by a short complementary “mask” oligo. This allows specific binding of the SNV probes
to the allelic RNA variants. The “mask” oligos are then displaced as the allele-specific oligos fully anneal with their target RNAs. Figure adapted from Levesque
et al. (2013)
Imprinted Genes and Hypothalamic Function
287
288 M. Pulix and A. Plagge
with tissue RNAseq analysis (see below) for determination of parental expression
biases.
Apart from these in situ techniques, molecular genetics approaches have been
established to analyze epigenetic features as well as RNA expression from single-
cell extracts (Angermueller et al. 2016; Cheow et al. 2015; Clark et al. 2018; Deng
et al. 2014; Kelsey et al. 2017). Some methods are designed for a low-cost, focused
investigation of specific loci (Cheow et al. 2015), while recent trends favor genome-
wide next-generation sequencing technologies. RNAseq was first applied by Deng
and colleagues (Deng et al. 2014) to investigate transcriptome-wide allelic expres-
sion in single cells using suitable strain-specific SNPs. Apart from parent-of-origin-
dependent (imprinted) gene expression, they observed that a substantial proportion
of genes (12–24%) are expressed monoallelically in a dynamic, random, and sto-
chastic way (Deng et al. 2014). The most recent technological advances enable a
combined analysis of DNA methylation, transcriptome, and nucleosome accessibil-
ity from single cells (scM&T-seq, scNMT-seq) (Angermueller et al. 2016; Kelsey
et al. 2017). These approaches allow direct correlations to be made between epige-
netic status, chromatin accessibility, and gene expression in a single cell. Although
we cannot cover these methods in detail here [for a comprehensive review see
Kelsey et al. (2017)], we would like to emphasize that consideration needs to be
given to the way single cells can be isolated and the overall number of cells required
for analysis. Most frequently, flow cytometry/fluorescence-activated cell sorting
(FACS) is used to obtain single cells or nuclei (Luo et al. 2017), although
microfluidics or manual collection via micromanipulators can also be applied
depending on the tissue or cell type of interest (Cheow et al. 2015). Furthermore,
and again depending on the complexity of the tissue to be investigated, usually
hundreds or thousands of cells need to be included in an experiment to obtain a
representative dataset for the cell population in question. As single-cell RNAseq and
methylome analyses have already been applied on some brain regions (Luo et al.
2017; Tasic et al. 2016; Zeisel et al. 2015), it will be exciting to see the multi-omics
approaches being adapted to neurobiological questions and specifically to character-
ize the large diversity of hypothalamic cell types in the future.
10.5 Perspectives
Within the brain, the hypothalamus has emerged as a major region in which imprinted
genes are expressed. Here, they exert crucial functions in the regulation of postnatal
development, energy homeostasis, hypothalamus–pituitary, and hypothalamus–thy-
roid neuroendocrine axes, as well as circadian rhythmicity and sleep. It is highly
likely that the list provided in this chapter will be expanded in the near future, as
several additional imprinted genes have been associated with a role in the hypothala-
mus, for example, Neuronatin, Mkrn3, Grb10, and Liz. On a molecular level, the
recent developments in single-cell techniques will shed further light on the epigenetic
regulation of imprinted genes. Many genes display an expression bias toward one
parental allele when analyzed on a whole-tissue level. To clarify whether this is due to
10 Imprinted Genes and Hypothalamic Function 289
a mixture of cell types, some of which showing biallelic others monoallelic expres-
sion, and to understand the neurobiological relevance of this, will be an exciting field
of future research.
Acknowledgment We would like to thank the Wellcome Trust PhD studentship program in
Cellular and Molecular Physiology at the University of Liverpool (099795/Z/12/Z) for research
support.
Key References
Buiting et al. (2016). This review provides an excellent overview about the clinical
symptoms and genetics of Angelman syndrome.
Cassidy et al. (2012). A recommended review summarising the clinical symptoms as
well as the genetics of Prader-Willi syndrome.
Chen et al. (2009). This paper shows for the first time genomic imprinting of Gnas in
the brain and that loss of maternal allele-specific expression of Gnas in the brain is
the cause for an obesity phenotype due to reduced energy expenditure.
Ferguson-Smith (2011) This review provides an excellent overview about the
history of genomic imprinting research and well-studied imprinted genes clusters.
Kelsey et al. (2017). This recent review gives a good introduction into novel single-
cell analysis techniques for epigenetics and gene expression.
Perez et al. (2015) This paper systematically analyzed imprinted gene expression
within the brain transcriptome with excellent detail.
Peters (2014). This review emphasizes the physiological functions of imprinted
genes.
Plagge et al. (2004). This paper identified the physiological functions of the pater-
nally expressed transcript variant Gnasxl of the Gnas locus.
References
Angermueller C, Clark SJ, Lee HJ, Macaulay IC, Teng MJ, Hu TX, Krueger F, Smallwood S,
Ponting CP, Voet T, Kelsey G, Stegle O, Reik W (2016) Parallel single-cell sequencing links
transcriptional and epigenetic heterogeneity. Nat Methods 13:229–232
Babak T, DeVeale B, Tsang EK, Zhou Y, Li X, Smith KS, Kukurba KR, Zhang R, Li JB, van der
Kooy D, Montgomery SB, Fraser HB (2015) Genetic conflict reflected in tissue-specific maps of
genomic imprinting in human and mouse. Nat Genet 47:544–549
Bartolomei MS, Ferguson-Smith AC (2011) Mammalian genomic imprinting. Cold Spring Harb
Perspect Biol 3:a002592
Bieth E, Eddiry S, Gaston V, Lorenzini F, Buffet A, Conte Auriol F, Molinas C, Cailley D,
Rooryck C, Arveiler B, Cavaille J, Salles JP, Tauber M (2015) Highly restricted deletion of
the SNORD116 region is implicated in Prader-Willi Syndrome. Eur J Hum Genet 23:252–255
Bischof JM, Stewart CL, Wevrick R (2007) Inactivation of the mouse Magel2 gene results in
growth abnormalities similar to Prader-Willi syndrome. Hum Mol Genet 16:2713–2719
Bonthuis PJ, Huang WC, Stacher Horndli CN, Ferris E, Cheng T, Gregg C (2015) Noncanonical
genomic imprinting effects in offspring. Cell Rep 12:979–991
290 M. Pulix and A. Plagge
presenting with severe pre- and post-natal growth retardation and intractable feeding difficulties.
Eur J Hum Genet 13:1033–1039
Gerard M, Hernandez L, Wevrick R, Stewart CL (1999) Disruption of the mouse necdin gene
results in early post-natal lethality. Nat Genet 23:199–202
Ginart P, Kalish JM, Jiang CL, Yu AC, Bartolomei MS, Raj A (2016) Visualizing allele-specific
expression in single cells reveals epigenetic mosaicism in an H19 loss-of-imprinting mutant.
Genes Dev 30:567–578
Haig D (2004) Genomic imprinting and kinship: how good is the evidence? Annu Rev Genet
38:553–585
Harris RA, Wang T, Coarfa C, Nagarajan RP, Hong C, Downey SL, Johnson BE, Fouse SD,
Delaney A, Zhao Y, Olshen A, Ballinger T, Zhou X, Forsberg KJ, Gu J, Echipare L, O’Geen H,
Lister R, Pelizzola M, Xi Y, Epstein CB, Bernstein BE, Hawkins RD, Ren B, Chung WY, Gu H,
Bock C, Gnirke A, Zhang MQ, Haussler D, Ecker JR, Li W, Farnham PJ, Waterland RA,
Meissner A, Marra MA, Hirst M, Milosavljevic A, Costello JF (2010) Comparison of
sequencing-based methods to profile DNA methylation and identification of monoallelic epige-
netic modifications. Nat Biotechnol 28:1097–1105
Hasegawa K, Kawahara T, Fujiwara K, Shimpuku M, Sasaki T, Kitamura T, Yoshikawa K (2012)
Necdin controls Foxo1 acetylation in hypothalamic arcuate neurons to modulate the thyroid
axis. J. Neurosci 32:5562–5572
He Q, Zhu Y, Corbin BA, Plagge A, Bastepe M (2015) The G protein alpha subunit variant XLαs
promotes inositol 1,4,5-trisphosphate signaling and mediates the renal actions of parathyroid
hormone in vivo. Sci Signal 8:ra84
Huang WC, Ferris E, Cheng T, Horndli CS, Gleason K, Tamminga C, Wagner JD, Boucher KM,
Christian JL, Gregg C (2017) Diverse non-genetic, allele-specific expression effects shape
genetic architecture at the cellular level in the mammalian brain. Neuron 93:1094–1109
Ivanova E, Kelsey G (2011) Imprinted genes and hypothalamic function. J Mol Endocrinol 47:
R67–R74
Kawashima S, Nakamura A, Inoue T, Matsubara K, Horikawa R, Wakui K, Takano K,
Fukushima Y, Tatematsu T, Mizuno S, Tsubaki J, Kure S, Matsubara Y, Ogata T, Fukami M,
Kagami M (2018) Maternal uniparental disomy for chromosome 20: physical and endocrino-
logical characteristics of five patients. J Clin Endocrinol Metab 103:2083–2088
Kelsey G, Stegle O, Reik W (2017) Single-cell epigenomics: recording the past and predicting the
future. Science 358:69–75
Kozlov SV, Bogenpohl JW, Howell MP, Wevrick R, Panda S, Hogenesch JB, Muglia LJ, Van
Gelder RN, Herzog ED, Stewart CL (2007) The imprinted gene Magel2 regulates normal
circadian output. Nat Genet 39:1266–1272
Krechowec SO, Burton KL, Newlaczyl AU, Nunn N, Vlatkovic N, Plagge A (2012) Postnatal
changes in the expression pattern of the imprinted signalling protein XLαs underlie the changing
phenotype of deficient mice. PLoS One 7:e29753
LaSalle JM, Reiter LT, Chamberlain SJ (2015) Epigenetic regulation of UBE3A and roles in human
neurodevelopmental disorders. Epigenomics 7:1213–1228
Lee S, Walker CL, Karten B, Kuny SL, Tennese AA, O’Neill MA, Wevrick R (2005) Essential role
for the Prader-Willi syndrome protein necdin in axonal outgrowth. Hum Mol Genet 14:627–637
Levesque MJ, Ginart P, Wei Y, Raj A (2013) Visualizing SNVs to quantify allele-specific
expression in single cells. Nat Methods 10:865–867
Li L, Keverne EB, Aparicio SA, Ishino F, Barton SC, Surani MA (1999) Regulation of maternal
behavior and offspring growth by paternally expressed Peg3. Science 284:330–333
Li YQ, Shrestha Y, Pandey M, Chen M, Kablan A, Gavrilova O, Offermanns S, Weinstein LS
(2016) G(q/11)α and G(s)α mediate distinct physiological responses to central melanocortins. J
Clin Invest 126:40–49
Luo C, Keown CL, Kurihara L, Zhou J, He Y, Li J, Castanon R, Lucero J, Nery JR, Sandoval JP,
Bui B, Sejnowski TJ, Harkins TT, Mukamel EA, Behrens MM, Ecker JR (2017) Single-cell
292 M. Pulix and A. Plagge
methylomes identify neuronal subtypes and regulatory elements in mammalian cortex. Science
357:600–604
Lyko F (2018) The DNA methyltransferase family: a versatile toolkit for epigenetic regulation. Nat
Rev Genet 19:81–92
Maillard J, Park S, Croizier S, Vanacker C, Cook JH, Prevot V, Tauber M, Bouret SG (2016) Loss
of Magel2 impairs the development of hypothalamic Anorexigenic circuits. Hum Mol Genet
25:3208–3215
Mantovani G, Bastepe M, Monk D, de Sanctis L, Thiele S, Usardi A, Ahmed SF, Bufo R,
Choplin T, De Filippo G, Devernois G, Eggermann T, Elli FM, Freson K, Garcia Ramirez A,
Germain-Lee EL, Groussin L, Hamdy N, Hanna P, Hiort O, Juppner H, Kamenicky P, Knight N,
Kottler ML, Le Norcy E, Lecumberri B, Levine MA, Makitie O, Martin R, Martos-Moreno GA,
Minagawa M, Murray P, Pereda A, Pignolo R, Rejnmark L, Rodado R, Rothenbuhler A,
Saraff V, Shoemaker AH, Shore EM, Silve C, Turan S, Woods P, Zillikens MC, Perez de
Nanclares G, Linglart A (2018) Diagnosis and management of pseudohypoparathyroidism and
related disorders: first international Consensus Statement. Nat Rev Endocrinol 14:476–500
Martinez ME, Charalambous M, Saferali A, Fiering S, Naumova AK, Germain DS, Ferguson-
Smith AC, Hernandez A (2014) Genomic imprinting variations in the mouse type 3 deiodinase
gene between tissues and brain regions. Mol Endocrinol 28:1875–1886
Matarazzo V, Caccialupi L, Schaller F, Shvarev Y, Kourdougli N, Bertoni A, Menuet C,
Voituron N, Deneris E, Gaspar P, Bezin L, Durbec P, Hilaire G, Muscatelli F (2017) Necdin
shapes serotonergic development and SERT activity modulating breathing in a mouse model for
Prader-Willi syndrome. elife 6:e32640
Mehta S, Williamson CM, Ball S, Tibbit C, Beechey C, Fray M, Peters J (2015) Transcription
driven somatic DNA methylation within the imprinted Gnas cluster. PLoS One 10:e0117378
Meister B, Perez-Manso M, Daraio T (2013) Delta-like 1 homologue is a hypothalamus- enriched
protein that is present in orexin-containing neurones of the lateral hypothalamic area.
J Neuroendocrinol 25:617–625
Mercer RE, Michaelson SD, Chee MJ, Atallah TA, Wevrick R, Colmers WF (2013) Magel2 is
required for leptin-mediated depolarization of POMC neurons in the hypothalamic arcuate
nucleus in mice. PLoS Genet 9:e1003207
Miller NL, Wevrick R, Mellon PL (2009) Necdin, a Prader-Willi syndrome candidate gene,
regulates gonadotropin-releasing hormone neurons during development. Hum Mol Genet
18:248–260
Mulchandani S, Bhoj EJ, Luo M, Powell-Hamilton N, Jenny K, Gripp KW, Elbracht M,
Eggermann T, Turner CL, Temple IK, Mackay DJ, Dubbs H, Stevenson DA, Slattery L, Zackai
EH, Spinner NB, Krantz ID, Conlin LK (2016) Maternal uniparental disomy of chromosome 20:
a novel imprinting disorder of growth failure. Genet Med 18:309–315
Muscatelli F, Abrous DN, Massacrier A, Boccaccio I, Le Moal M, Cau P, Cremer H (2000)
Disruption of the mouse Necdin gene results in hypothalamic and behavioral alterations
reminiscent of the human Prader-Willi syndrome. Hum Mol Genet 9:3101–3110
Nunn N, Feetham CH, Martin J, Barrett-Jolley R, Plagge A (2013) Elevated blood pressure, heart
rate and body temperature in mice lacking the XLαs protein of the Gnas locus is due to increased
sympathetic tone. Exp Physiol 98:1432–1445
Olova N, Krueger F, Andrews S, Oxley D, Berrens RV, Branco MR, Reik W (2018) Comparison of
whole-genome bisulfite sequencing library preparation strategies identifies sources of biases
affecting DNA methylation data. Genome Biol 19:33
Perez JD, Rubinstein ND, Fernandez DE, Santoro SW, Needleman LA, Ho-Shing O, Choi JJ,
Zirlinger M, Chen SK, Liu JS, Dulac C (2015) Quantitative and functional interrogation of
parent-of-origin allelic expression biases in the brain. elife 4:e07860
Peters J (2014) The role of genomic imprinting in biology and disease: an expanding view. Nat Rev
Genet 15:517–530
Plagge A (2012) Non-coding RNAs at the Gnas and Snrpn-Ube3a imprinted gene loci and their
involvement in hereditary disorders. Front Genet 3:264
10 Imprinted Genes and Hypothalamic Function 293
Plagge A, Gordon E, Dean W, Boiani R, Cinti S, Peters J, Kelsey G (2004) The imprinted signaling
protein XLαs is required for postnatal adaptation to feeding. Nat Genet 36:818–826
Plagge A, Isles AR, Gordon E, Humby T, Dean W, Gritsch S, Fischer-Colbrie R, Wilkinson LS,
Kelsey G (2005) Imprinted Nesp55 influences behavioral reactivity to novel environments. Mol
Cell Biol 25:3019–3026
Plasschaert RN, Bartolomei MS (2014) Genomic imprinting in development, growth, behavior and
stem cells. Development 141:1805–1813
Podyma B, Sun H, Wilson EA, Carlson B, Pritikin E, Gavrilova O, Weinstein LS, Chen M (2018)
The stimulatory G protein Gsα is required in melanocortin 4 receptor-expressing cells for
normal energy balance, thermogenesis, and glucose metabolism. J Biol Chem
293:10993–11005
Polex-Wolf J, Lam BY, Larder R, Tadross J, Rimmington D, Bosch F, Cenzano VJ, Ayuso E, Ma
MK, Rainbow K, Coll AP, O’Rahilly S, Yeo GS (2018) Hypothalamic loss of Snord116
recapitulates the hyperphagia of Prader-Willi syndrome. J Clin Invest 128:960–969
Richard N, Molin A, Coudray N, Rault-Guillaume P, Juppner H, Kottler ML (2013) Paternal GNAS
mutations lead to severe intrauterine growth retardation (IUGR) and provide evidence for a role
of XLαs in fetal development. J Clin Endocrinol Metab 98:E1549–E1556
Schaller F, Watrin F, Sturny R, Massacrier A, Szepetowski P, Muscatelli F (2010) A single
postnatal injection of oxytocin rescues the lethal feeding behaviour in mouse newborns deficient
for the imprinted Magel2 gene. Hum Mol Genet 19:4895–4905
Shi SQ, Bichell TJ, Ihrie RA, Johnson CH (2015) Ube3a imprinting impairs circadian robustness in
Angelman syndrome models. Curr Biol 25:537–545
Skryabin BV, Gubar LV, Seeger B, Pfeiffer J, Handel S, Robeck T, Karpova E, Rozhdestvensky
TS, Brosius J (2007) Deletion of the MBII-85 snoRNA gene cluster in mice results in postnatal
growth retardation. PLoS Genet 3:e235
Stohn JP, Martinez ME, Zafer M, Lopez-Espindola D, Keyes LM, Hernandez A (2018) Increased
aggression and lack of maternal behavior in Dio3-deficient mice are associated with
abnormalities in oxytocin and vasopressin systems. Genes Brain Behav 17:23–35
Tacer KF, Potts PR (2017) Cellular and disease functions of the Prader-Willi Syndrome gene
MAGEL2. Biochem J 474:2177–2190
Tasic B, Menon V, Nguyen TN, Kim TK, Jarsky T, Yao Z, Levi B, Gray LT, Sorensen SA,
Dolbeare T, Bertagnolli D, Goldy J, Shapovalova N, Parry S, Lee C, Smith K, Bernard A,
Madisen L, Sunkin SM, Hawrylycz M, Koch C, Zeng H (2016) Adult mouse cortical cell
taxonomy revealed by single cell transcriptomics. Nat Neurosci 19:335–346
Tommiska J, Kansakoski J, Skibsbye L, Vaaralahti K, Liu X, Lodge EJ, Tang C, Yuan L,
Fagerholm R, Kanters JK, Lahermo P, Kaunisto M, Keski-Filppula R, Vuoristo S, Pulli K,
Ebeling T, Valanne L, Sankila EM, Kivirikko S, Laaperi M, Casoni F, Giacobini P, Phan-
Hug F, Buki T, Tena-Sempere M, Pitteloud N, Veijola R, Lipsanen-Nyman M, Kaunisto K,
Mollard P, Andoniadou CL, Hirsch JA, Varjosalo M, Jespersen T, Raivio T (2017) Two
missense mutations in KCNQ1 cause pituitary hormone deficiency and maternally inherited
gingival fibromatosis. Nat Commun 8:1289
Turan S, Bastepe M (2013) The GNAS complex locus and human diseases associated with loss-of-
function mutations or epimutations within this imprinted gene. Horm Res Paediatr 80:229–241
Villanueva C, Jacquier S, de Roux N (2012) DLK1 is a somato-dendritic protein expressed in
hypothalamic arginine-vasopressin and oxytocin neurons. PLoS One 7:e36134
Weinstein LS (2014) Role of G(s)α in central regulation of energy and glucose metabolism. Horm
Metab Res 46:841–844
Wijesuriya TM, De Ceuninck L, Masschaele D, Sanderson MR, Carias KV, Tavernier J, Wevrick R
(2017) The Prader-Willi syndrome proteins MAGEL2 and necdin regulate leptin receptor cell
surface abundance through ubiquitination pathways. Hum Mol Genet 26:4215–4230
Wu Z, Martinez ME, St Germain DL, Hernandez A (2017) Type 3 deiodinase role on central
thyroid hormone action affects the leptin-melanocortin system and circadian activity. Endocri-
nology 158:419–430
294 M. Pulix and A. Plagge
Xie T, Plagge A, Gavrilova O, Pack S, Jou W, Lai EW, Frontera M, Kelsey G, Weinstein LS (2006)
The alternative stimulatory G protein α-subunit XLαs is a critical regulator of energy and
glucose metabolism and sympathetic nerve activity in adult mice. J Biol Chem
281:18989–18999
Zeisel A, Munoz-Manchado AB, Codeluppi S, Lonnerberg P, La Manno G, Jureus A, Marques S,
Munguba H, He L, Betsholtz C, Rolny C, Castelo-Branco G, Hjerling-Leffler J, Linnarsson S
(2015) Brain structure. Cell types in the mouse cortex and hippocampus revealed by single-cell
RNA-seq. Science 347:1138–1142
Rhythmic Epigenetics in Neuroendocrine
and Immune Systems 11
Christopher S. Coyle, Elisabetta Tolla, and Tyler J. Stevenson
Abstract
Biological rhythms in neuroendocrine and immune systems are pervasive. Daily
and seasonal changes in day lengths regulate multiple physiological and immuno-
logical parameters in a diverse range of animals. A series of studies have shown
that epigenetic modifications exhibit naturally occurring rhythms across short- and
long-term timescales. In this chapter, we describe daily, estrous and seasonal
oscillations in epigenetic enzymes in neuroendocrine substrates, peripheral repro-
ductive tissues and immune cells (e.g. leukocytes). The predominant focus of
the chapter includes enzymes involved in DNA methylation and histone
modifications, such as DNA methyltransferase and histone deacetylases. The
findings presented herein highlight that epigenetic modifications can be permanent
as well as transient with long-term consequences on the timing of physiological
and behavioural processes. Moreover, the bidirectional interaction between the
immune system and the neuroendocrine nucleus that controls biological rhyth-
micity, the suprachiasmatic nucleus, emphasizes the need to understand rhythmic
changes in epigenetic enzymes and the consequences of disrupted daily and
seasonal rhythms.
Keywords
Methyltransferase · Acetylation · Oscillation · Circadian · Circannual
11.1 Introduction
The study of epigenetics originates in the field of developmental biology. Early work
by Conrad Waddington synthesized the concept that interactions between the envi-
ronment and the genetic material during embryogenesis and early life experiences
shaped the mature, complex organism (Waddington 1953). As the field advanced, it
became clear that experience-dependent modifications could alter an individual’s
fitness, and this was subsequently passed to the next generation; commonly referred
to as transgenerational inheritance. Here, the study of mitotically and/or meiotically
heritable changes in gene function was observed independent of changes in the
genome sequence (Riggs and Porter 1996). In parallel, the identities of enzymes
that act to catalyze the biochemical modifications of genome function (i.e. DNA
methylation and histone modification) were discovered to include: histone (de)
acetylases (Taunton et al. 1996) and DNA methyltransferases (Yoder et al. 1997).
Today, the next level of complexity focuses on RNA methylation and non-coding
RNA mechanisms (Guil and Esteller 2009). For the past 20 years, there has been an
exponential growth in our understanding of the environmental and hormonal regula-
tion of epigenetic enzymes (Stevenson 2017a), particularly in the field of cancer
genomics (Esteller 2008) with significantly less attention in the field of
11 Rhythmic Epigenetics in Neuroendocrine and Immune Systems 297
Epigenetic changes that impact the genome (i.e. DNA methylation) and chromatin
(i.e. histone acetylation) are induced by a range of enzyme families that show tissue-
specific expression profiles. DNA methylation on the genome template is regulated
by DNA methyltransferases (DNMT) and Ten-Eleven-Translocation methylcytosine
dioxygenase (TET) enzymes (Fig. 11.1a). DNMTs catalyze the removal of methyl
from S-adenosyl methionine (SAM) and preferentially add to cytosine-guanine
(CpG) residues, but also non-CpG residues (Jang et al. 2017). Conversely, TET
enzymes oxidize 5-methylcytosine that is then actively reverted to cytosine through
oxidation and thymine DNA glycosylase (TDG)-mediated base excision repair
(Kohli and Zhang 2013). Both DNMT and TET enzymes have a number of different
isoforms that have distinct structure–function relations that contribute to the devel-
opmental- and cell-specificity of epigenetic modifications. In adults, DNMT1, 3a, and
298 C. S. Coyle et al.
A SAM SAH
B H2O Acetate
DNMT HDAC
euchromatin heterochromatin
TET HAT
Fig. 11.1 Functional pathways of epigenetic enzymes. (a) DNA methyltransferase (DNMT)
enzymes catalyze the addition of methyl groups, depicted by red dots, on to the genome template.
Ten Eleven Translocation (TET) enzymes initiate a cascade of biochemical changes that include the
oxidation of 5-methylcytosine and promote locus-specific reversal of DNA methylation indicated
by grey dots. (b) Epigenetic enzymes involved in chromatin modifications include histone
deacetylase (HDAC) and histone acetylase (HAT). HDAC triggers the removal of acetyl groups
from histone. HAT enzymes catalyze the removal of acetyl groups from Acetyl-CoA and add acetyl,
indicated by red curvilinear lines, on to histones. Abbreviations: S-adenosyl methionine (SAM), S-
adenosyl homocysteine (SAH), carbon dioxide (CO2), alpha-ketoglutarate (αKG), coenzyme A
(CoA)
3b exhibit wide anatomical distribution with high levels of expression in the brain,
reproductive tissues, and immune glands (e.g. spleen) (Fig. 11.2). The TET1-3
enzymes also show high expression levels in the brain and female reproductive
tissues (i.e. ovary) (Fig. 11.2).
Chromatin modifications include histone acetylation induced by histone
acetylases (HAT/KAT) and the converse histone deacetylase (HDAC) enzymes
that consist of several isoforms (Klose and Bird 2006). HAT enzymes catalyze the
removal of acetyl from acetyl-coenzyme A (CoA) and induce the transformation from
closed heterochromatin into the transcriptional accessible euchromatin states
(Fig. 11.1b). The removal of acetyl groups from histones by HDACs results in
silenced gene transcription and the formation of cellular acetate. Similar to the
enzymes involved in DNA methylation, both HDAC and HAT enzymes show higher
levels of expression in the brain, reproductive tissues and immune glands (Fig. 11.2).
The initial evidence to indicate that epigenetic enzymes display oscillations were
derived from high-throughput cDNA microarrays and RNA sequencing platforms
(Stevenson 2018). The current evidence suggests that daily epigenetic enzyme
oscillations are isoform-specific and tissue-dependent. For example, in the adult
11 Rhythmic Epigenetics in Neuroendocrine and Immune Systems 299
Lung Brain
DNMT3a TET2 DNMT1 TET1
HDAC3 TET3 DNMT3a TET2
KAT2a DNMT3b TET3
HDAC1 HAT1
Heart HDAC2 KAT2a
DNMT3a TET1 HDAC3 KAT2b
HDAC3 KAT2b
Spleen
Lymph DNMT1 TET2
DNMT1 HAT1 DNMT3a TET3
DNMT3a DNMT3b HAT1
HDAC1 HDAC1 KAT2a
HDAC3 HDAC3 KAT2b
Liver Ovary
DNMT1 HAT1 DNMT1 TET1
DNMT3b KAT2b DNMT3a TET2
HDAC1 DNMT3b TET3
HDAC2 HDAC2 KAT2a
HDAC3 KAT2b
Adipose
TET1 Uterus
TET2 DNMT1
KAT2a DNMT3a
KAT2b DNMT3b
HDAC1
HDAC2
Testis HDAC3
DNMT1 HAT1
DNMT3a KAT2a
DNMT3b Placenta
HDAC1 DNMT1 TET1
HDAC2 DNMT3a TET2
HDAC3 DNMT3b TET3
HDAC1 HAT1
HDAC2 KAT2a
HDAC3
Fig. 11.2 Distribution and relative abundance of epigenetic enzymes. A representative diagram to
highlight the tissue-specific and expression levels of epigenetic enzymes in multiple tissues.
Epigenetic enzymes presented in tissue boxes were selected based on higher than average Reads
Per Kilobase Million (RPKM) expression levels across all tissues. Most enzymes involved in DNA
methylation and histone modifications are expressed in neuroendocrine substrates (i.e. brain) as
well as reproductive tissues (i.e. testes, ovary and uterus). RPKM data were obtained from
PUBMED and are based on Fagerberg et al. (2014)
300 C. S. Coyle et al.
mouse suprachiasmatic nucleus (SCN), the master neuroendocrine nucleus for the
regulation of daily rhythms (Hastings et al. 2014), DNMT3a is expressed, exhibits
daily rhythms in mRNA levels and serves to maintain the integrity of circadian
oscillations in genome function (Azzi et al. 2014). In the adult male Siberian hamster
there are robust daily oscillations in hypothalamic dnmt3a and hdac4 expression
with peak levels during the dark phase that is followed by a precipitous decline after
lights on (Stevenson 2017b). However, not all epigenetic enzymes exhibit daily
variation in neuroendocrine expression, and include DNMT1, DNMT3b, HDAC1,
HDAC2 and HDAC3. The daily variation in epigenetic enzyme expression
(e.g. DNMT3a) is probably driven by the circadian clock gene aryl hydrocarbon
receptor nuclear translocator-like 1 (BMAL1) proteins as there are 3 BMAL1
binding sites in the DNMT3a promoter and there is robust daily variation in
BMAL1 binding in the DNMT3a promoter (Fig. 11.3a) (Koike et al. 2012).
In addition to daily rhythms, epigenetic enzymes have been shown to exhibit
robust variation across annual scales. Most experimental evidence shows that neu-
roendocrine epigenomic plasticity is a major contributor to the seasonal phenotypic
variation in mammalian and avian species. In the Siberian hamster brain, there is a
significantly higher level of global DNA methylation in the hypothalamus during the
summer breeding conditions compared to the winter non-breeding conditions
(Stevenson and Prendergast 2013). The decrease in global DNA methylation occurs
in parallel to reduced daily dnmt1 and dnmt3a levels (Fig. 11.3b) (Stevenson 2017b).
These data indicate that reduced hypothalamic DNA methylation may be an endog-
enous circannual mechanism for the timing of seasonal life–history transitions
(Helm and Stevenson 2015; Stevenson and Lincoln 2017). There are also significant
seasonal oscillations in DNMT expression in hamster reproductive tissues. During
the breeding period, DNMT3a levels are significantly reduced in the testes and
uterine tissue (Lynch et al. 2016) (Fig. 11.3c), liver (Alvarado et al. 2015), skeletal
muscle (Alvarado et al. 2015) and circulating leukocytes (Stevenson et al. 2014). In
humans, there are also marked seasonal changes in DNA methylation. Blood
samples revealed that in spring and summer months, there were significantly higher
levels of DNA methylation in the promoters for cancer-related genes (RASSF1A and
MGMT) compared to autumn and winter months (Ricceri et al. 2014). The exact
mechanisms for how external information, such as day length, regulates epigenetic
enzyme expression are not well known. In addition to day length, supplementary
cues including temperature and diet (see below), likely contribute to the regulation of
epigenetic enzyme expression in a tissue-dependent manner. The two primary
hormonal candidates that provide an internal code of environmental cues are mela-
tonin and gonadal-derived oestrogen (Stevenson 2017a). In human embryonic
kidney cells (HEK293), the administration of melatonin induced an increase in
dnmt3a expression (Fig. 11.3d). Given that melatonin is a reliable physiological
measurement of ‘night-length’ a proxy for night duration, these data suggest that the
annual change in day length drives melatonin-dependent entrainment of epigenetic
enzyme expression, and DNMT3a in particular. In ovariectomized hamsters, a bolus
injection of estradiol and progesterone (E2P4) significantly reduced dnmt3a expres-
sion (Fig. 11.3e). It is likely that oestrogen actions provide supplementary regulation
11 Rhythmic Epigenetics in Neuroendocrine and Immune Systems 301
A)
DNMT3a promoter
1A 1B 1C 2
DNMT3a exons
B) C)
10 4
LD
SD
dnmt3a expression
mRNA expression
8
3
6
2
1
2
0 0
dnmt1 dnmt3a dnmt3b Testes Ovary Uterus
D) E)
4 6
5
dnmt3a expression
dnmt3a expression
3
4
2 3
2
1
1
0 0
Saline 1nM 10nM 100nM Oil 12hr 24hr
Fig. 11.3 Photoperiodic and hormonal regulation of DNA methyltransferase 3a. DNMT3a is one
epigenetic enzyme that exhibits consistent rhythmic expression in multiple nuclei and tissues. (a)
The dnmt3a promoter has several binding motifs for the circadian clock gene BMAL1. Koike et al.
(2012) confirmed that BMAL1 directly binds to transcriptional regulatory regions for dnmt3a and
shows daily changes with peak activity during the late dark phase. (b) Hypothalamic dnmt1 and
dnmt3a expressions display photoperiod-dependent regulation in which long summer-like day
lengths (LD) stimulate higher levels and short winter-like days (SD) reduce levels. (c) Conversely,
there is an increase in dnmt3a expression in the testes and uterine tissue of SD reproductively
regressed hamsters. The annual change in photoperiod is encoded by high amplitude changes in the
duration of melatonin and plasma oestrogen concentrations. (d) Melatonin was sufficient to induce
dnmt3a expression in HEK293 cells. (e) A single bolus injection of E2P4 to ovariectomized
hamsters significantly reduced uterine dnmt3a expression. Data are adapted from Stevenson
(2017a, b) and Lynch et al. (2016)
302 C. S. Coyle et al.
Uterus
Morphology
Progesterone
Hormone
Estradiol
HDAC2
Epigenetic
DNMT3b
DNMT3a
Fig. 11.4 Epigenetic enzymes oscillate in uterine tissues across the estrous cycle. The female
rodent estrous cycle is characterized by robust changes in morphology, plasma hormones and
uterine epigenetic enzymes. (Upper Panel) During the estrous cycle, uterine mass significantly
increases during the transition from proestrous to estrous stages. On the evening of proestrous, there
is a significant increase in the probability and intensity of female reproductive behaviours. The
green bar depicts ‘behavioural estrous’, the period of high proceptivity and receptivity in females.
(Middle Panel) Cycling changes in ovarian hormones estradiol and progesterone, the primary
hormones that induce marked morphological changes in the uterus and transcriptional activation
of several genes involved in the control of reproductive physiology. (Lower Panel) Three epigenetic
enzymes in the uterus, HDAC2, DNMT3a and DNMT3b, were found to be inhibited by estradiol/
progesterone. Data are adapted from Lynch et al. (2016, 2017)
304 C. S. Coyle et al.
There are multiple systems that control energy intake and expenditure. A complex
interaction of neuroendocrine substrates, adipose tissue, muscle cells, pancreatic
cells and liver cells serve to maintain a homeostatic steady-state condition (Yeo
and Heisler 2012). Daily rhythms in epigenetic enzymes have been characterized in
discrete hypothalamic nuclei and peripheral tissues that suggest a complex
epigenomic reprogramming occurs daily to ensure optimal energy balance (Asher
and Sassone-Corsi 2015). The development of a freely available, online resources
that contain massive datasets based on high-throughput methods (e.g. microarray,
RNA sequencing), such as CircaDB (circadb.hogeneschlab.org), provide a valuable
opportunity to interrogate daily rhythms in epigenetic enzyme RNA expression
across multiple tissues. For example, CircaDB analyzes reveal that brown and
white adipose tissues show daily rhythms in hdac6 and kat2b; which coincidentally
have anti-phase expression waveforms (Pizarro et al. 2013). hdac6 expression
peaked during the mid-point of a 12:12 LD cycle, whereas kat2b peaks during the
dark phase. The enzyme expression patterns here suggest a daily switch in histone
(de)acetylation occurs that is not driven by food intake and instead, reflects an
endogenous timing mechanism. In the liver, tet2 and dnmt3b peak prior to lights
on; which suggest the potential for daily variation in epigenetic tone; high levels of
both methylation and demethylation prior to lights on followed by a subsequent
decline in epigenomic activity during the dark phase. However, other research has
shown that liver dnmt3a peaks during the mid-point of the light phase (Hughes et al.
2009). Conversely, liver hdac5 peaks during the dark phase, and hdac6 peaks during
the light phase.
There is growing evidence that the time of food intake has a remarkable level of
bidirectional interactions between the circadian clock and energy balance (Asher and
Sassone-Corsi 2015). The vast majority of research has examined the contribution of
chromatin remodelling via the activity of sirtuin proteins (SIRT1) within well-
defined hypothalamic nuclei involved in the regulation of metabolism, appetite
and food intake (Asher and Sassone-Corsi 2015). SIRT1 removes histone acetyl
groups via a nicotinamide adenine dinucleotide (NAD+)-dependent manner.
Increased SIRT1 activity in the ventromedial nucleus (VMN) of the hypothalamus
11 Rhythmic Epigenetics in Neuroendocrine and Immune Systems 305
leads to histone deacetylation and can directly interact with the circadian clock gene
CLOCK; and thereby, regulates clock-controlled genes in a histone deacetylation-
dependent manner (Nakahata et al. 2008; Orozco-Solis et al. 2015). The current
evidence suggests that SIRT1 expression specifically within VMN Sf1 neurons is
essential to coordinate circadian food intake and daily energetic state.
Immune responses are dynamically regulated over daily and seasonal timescales
(Stevenson and Prendergast 2015). This section will give an overview of the ongoing
research regarding the role of epigenetic mechanisms in immune function, focusing
on evidence that indicates reversible modifications. The epigenetic enzymes
described above have been implicated in pathological conditions and progression,
e.g. cancer, and have become a powerful tool in disease treatment (Widschwendter
et al. 2018). One tantalizing conjecture is the potential loss of natural rhythmic
epigenetic modifications as a contributing factor to disease progression. Therefore,
the investigation of natural rhythmic variation in epigenetic modifications can give
clinical researchers insight into the factors that could be involved in enzyme
disruption, as well as aiding in the development of novel reversible treatment
strategies.
It is becoming clear that immune function (e.g. cancer) and biological rhythms
(i.e. circadian) have bidirectional interactions. The increased attention to disrupted
circadian rhythms has resulted in the identification that entrained daily rhythms
contribute to health and well-being (Savvidis and Koutsilieris 2013). Leukocyte
migration from lymph nodes, lymph and plasma shows daily rhythms with peak
CD4+, CD8+ T cells and B cell levels during the light phase in mice (Druzd et al.
2017) and Siberian hamsters (Prendergast et al. 2013). Disruption of circadian
timing induces leukocyte arrhythmia, reduced daily amplitudes in circadian clock
gene and inhibited CD11+ dendritic cells, indicating that a functional SCN is
required for daily variation in immune function (Prendergast et al. 2013). Likewise,
there are distinct seasonal rhythms in plasma leukocytes numbers and leukocyte
transcriptomic profiles in hamsters and humans, respectively (Bilbo et al. 2002;
Banks et al. 2016; Dopico et al. 2015). During winter photoperiods, there are
significantly elevated circulating blood leukocytes, T cells and natural killer cells
(Bilbo et al. 2002). The winter enhancement of immune function and survival is
common across species and serves to increase survival (Nelson 2004). The increase
in many chronic diseases and pathological states has been attributed to not only
disrupted circadian rhythms as described above, but also disruption in seasonal
immune rhythms (Stevenson et al. 2015).
A few studies have examined the daily or seasonal variation of epigenetic enzyme
expression in immune tissues or cells. A massive transcriptomic analysis of seasonal
waveforms of human leukocytes identified that dnmt1, hdac5, -6 and -9 vary across
the year (Dopico et al. 2015). In Siberian hamsters, short photoperiods resulted in
higher dnmt3b leukocyte expression (Stevenson et al. 2014). The increased dnmt3b
306 C. S. Coyle et al.
activity within the immune response has been explored both in vitro and in vivo
(Okano et al. 1999; Fitzpatrick and Wilson 2003). Similar to DNMT1, their expres-
sion is required for normal development and cell survival, as their loss results in
embryonic lethality (Okano et al. 1999).
It is clear that both genetics and epigenetics play a role in the development of the
majority of neuroendocrine cancers. One the ‘10 hallmarks of cancer’ outlined by
Hanahan and Weinberg (2011) is genomic instability that results from alterations in
the epigenome (Esteller 2008; Hanahan and Weinberg 2011), including both global
hypomethylation (Gaudet et al. 2003) and gene-specific hypermethylation (Herman
and Baylin 2003). High expression of dnmts has been shown in a variety of cancers,
e.g. neuroblastoma, prostate cancer (Kautiainen and Jones 1986). Several studies
have assessed the function of DNMT1 through both pharmacological inhibition and
gene knockdown in vitro and in vivo. DNMT1 is required for development in mice,
as dnmt1 / is lethal (Li et al. 1992). This has been attributed to the enzyme’s
essential function in maintaining DNA methylation patterns in daughter cells
through generations (Lei et al. 1996), a function necessary for normal survival and
growth. Jackson-Grusby et al. (2001) observed that mice dnmt1 / embryonic stem
cell-derived fibroblasts undergo a p53-dependent cell death. p53 is a tumour sup-
pressor gene, involved in inhibiting cell cycle progression when errors are detected
in the genome (Symonds et al. 1994). When methylated, p53 expression is silenced
and results in survival of cancerous cells (Jackson-Grusby et al. 2001). Knox et al.
(2000) monitored the effects of dnmt1 / in human lung cancer cells and found that
hypomethylation due to the absence of dnmt1 also inhibits cell proliferation through
apoptosis. The transcription of other tumour suppressor genes, such as p16 and p21,
was found to be upregulated after using pharmacological inhibitors of DNMT1
enzymatic activity in the human bladder (Fournel et al. 1999) and colon (Rhee
et al. 2000) carcinoma cells, preventing tumour progression. It is intriguing that both
expression of tumour suppressor genes (Miki et al. 2013) and epigenetic enzymes
has been shown to vary in a circadian manner. p53 is able to regulate the cell cycle is
via ‘clocking’ genes, by inhibiting the per2 gene through interaction with BMAL1
and CLOCK (Miki et al. 2013; Fu et al. 2002). Repression of per2 expression has
been associated with overexpression of c-MYC proto-oncogene (cMyc), a gene that
codes for transcription factors involved in cell cycle progression (Pelengaris et al.
2002). cMyc directly binds to DNMT3a to promote cell growth (Brenner et al.
2005). It is then possible that circadian rhythms dictate patterns of certain epigenetic
components, resulting in cancer development. However, the exact mechanisms
underlying the circadian control of epigenetic signals in tumourigenesis are still
largely unknown.
Aberrant expression of HAT and HDAC activity also contributes to the rise and
progression of specific diseases (Jones and Baylin 2007) and include cancer (Cress
and Seto 2000), Alzheimer’s disease (Francis et al. 2009) and mental retardation
(Coffee et al. 1999). HDAC activity has been shown to be linked to the expression of
certain eukaryotic cell cycle regulation factors, such as retinoblastoma (RB) protein
(Brehm et al. 1998). Disrupted interaction between HDAC proteins and RB proteins
prevents the apoptosis of cancerous cells, allowing tumourigenesis (Xu et al. 1993).
308 C. S. Coyle et al.
It has been suggested that an ‘epigenetic homeostasis’ exists and is needed for
normal cellular function that is characterized by a balance between DNA (de)
methylation as well as histone (de)acetylation (Jones and Baylin 2007). In this
chapter, we have focused on ‘epigenetic plasticity’, the transient modifications to
the genome and/or chromatin template that leads to a functional change in the
neuroendocrine control of reproduction, energy balance and immune function. The
predominant difference between these philosophical positions is that the disruption
of epigenetic homeostasis leads to long-term consequences in cellular function, such
as chronic diseases while the epigenetic plasticity model argues that the short- and
long-term oscillations in epigenetic enzymes maintain cyclical cell- and tissue-
function that is required for naturally occurring, normal daily, estrous and seasonal
rhythms (Stevenson 2018). A greater understanding of the mechanistic and func-
tional basis of the latter can yield significant insights into the genetic control of
animal health and wellness. Below, we outlined two major perspectives for future
research.
The field of epigenetic research is exciting as ongoing findings reveal the massive
potential for basic, translational and applied research. The advent of high-throughput
sequencing platforms such as Whole-Genome Sequencing (WGS), transcriptomic,
proteomic assays has yielded immense information on the complexity of molecular
and cellular signalling pathways in neuroendocrine and immune cells. Moreover, the
chemical modification of genome structure using sodium bisulfite now provides the
capability to identify specific nucleotide residues/motifs that show variation in
epigenetic modifications (i.e. methylation). This chapter has focused on the rhythmic
expression of multiple enzymes involved in epigenetic modifications. The data show
tissue-, cell- and time-dependent oscillations; however, our understanding of the
downstream effects is limited. The next steps are to establish the precise genomic
and chromatin location of daily, estrous and seasonal epigenetic modifications. The
combination of sequencing platforms, such as Illumina or PacBIO, provides a
valuable opportunity to address wherein the genome epigenetic modifications
exhibit rhythmic variation; and thus, novel gene regulatory mechanisms within
tissues and more importantly at a single-cell resolution.
Another advantage of sequencing tools to identify epigenetic modifications is the
development of individual or personalized approaches to medical treatments (Rasool
et al. 2015). Indeed, further investigation into the exact genomic and chromatin
11 Rhythmic Epigenetics in Neuroendocrine and Immune Systems 309
Key References
Azzi et al. (2014). This study shows that daily oscillations in SCN DNA methylation
regulates circadian clock function.
Bilbo et al. (2002). This is an exciting report on the seasonal regulation of immune
function.
Dopico et al. (2015). This study reveals large annual changes in leukocyte
transcriptomes in humans.
Koike et al. (2012). One key finding identified in this paper were daily rhythms in
BMAL1 binding on the dnmt3a promoter.
Lynch et al. (2016). This report illustrated increased dnmt3a expression was nega-
tively associated with oscillations in fertility.
Nugent et al. (2015). This study discovered sex-differences in neuroendocrine
DNMT3a and a lack of expression facilitated masculinization.
Stevenson and Prendergast. (2013). This study established that rhythmic neuroen-
docrine DNA methylation underlies seasonal timing of reproduction.
References
Agarwal S, Rao A (1998) Modulation of chromatin structure regulates cytokine gene expression
during T cell differentiation. Immunity 9:765–775
Ahluwalia A, Hurteau JA, Bigsby RM, Nephew KP (2001) DNA methylation in ovarian cancer.
II. Expression of DNA methyltransferases in ovarian cancer cell lines and normal ovarian
epithelial cells. Gynecol Oncol 82:299–304
Alvarado S, Mak T, Liu S, Storey KB, Szyf M (2015) Dynamic changes in global and gene-specific
DNA methylation during hibernation in adult thirteen-lined ground squirrels, Ictidomys
tridecemlineatus. J Exp Biol 218:1787–1795
Asher G, Sassone-Corsi P (2015) Time for food: the intimate interplay between nutrition, metabo-
lism and the circadian clock. Cell 161:84–92
Azzi A, Dallmann R, Casserly A, Rehrauer RH, Patrignani A, Maier B, Kramer A, Brown SA
(2014) Circadian behavior is light-programmed by plastic DNA methylation. Nat Neurosci
17:377–382
Banks R, Delibegovic M, Stevenson TJ (2016) Photoperiod- and triiodothyronine-dependent
regulation of reproductive neuropeptides, proinflammatory cytokines, and peripheral physiol-
ogy in Siberian hamsters (Phodopus sungorus). J Biol Rhythm 31:299–307
Bilbo SD, Dhabhar FS, Viswanathan K, Saul A, Yellon SM, Nelson RJ (2002) Short day lengths
augment stress-induced leukocyte trafficking and stress-induced enhancement of skin immune
function. Proc Natl Acad Sci U S A 99:4067–4072
310 C. S. Coyle et al.
Fu L, Pelicano H, Liu J, Huang P, Lee CC (2002) The circadian gene Period2 plays an important
role in tumor suppression and DNA damage response in vivo. Cell 111:41–50
Gaudet F, Hodgson JG, Eden A, Jackson-Grusby L, Dausman J, Gray JW, Leonhardt J, Jaenisch R
(2003) Induction of tumors in mice by genomic hypomethylation. Science 300:489–492
Greives TJ, Kriegsfeld LJ, Bentley GE, Tsutsui K, Demas GE (2008) Recent advances in reproduc-
tive neuroendocrinology: a role for RFamide peptides in seasonal reproduction? Proc Biol Sci
275:1943–1951
Guil S, Esteller M (2009) DNA methylomes, histone codes and miRNAs: tying it all together. Int J
Biochem Cell Biol 41:87–95
Hanahan D, Weinberg R (2011) Hallmarks of cancer: the next generation. Cell 144:646–674
Hastings MH, Brancaccio M, Maywood ES (2014) Circadian pacemaking in cells and circuits of the
suprachiasmatic nucleus. J Neuroendocrinol 26:2–10
Heard E, Martienssen RA (2014) Transgenerational epigenetic inheritance: myths and mechanisms.
Cell 157:95–109
Helm B, Stevenson TJ (2015) Circannual rhythms: history, present challenges, future directions. In:
Numata H, Helm B (eds) Annual, lunar and tidal clocks: patterns and mechanisms of nature’s
enigmatic rhythms. Springer Japan, Tokyo, pp 213–225
Herman JG, Baylin SB (2003) Gene silencing in cancer in association with promoter
hypermethylation. N Engl J Med 349:2042–2054
Hughes ME, DiTacchio L, Hayes KR, Vollmers C, Pulivarthy S, Baggs JE, Panda S, Hogenesch JB
(2009) Harmonics of circadian gene transcription in mammas. PLoS Genet 5:e1000442
Jackson-Grusby L, Beard C, Possemato R, Tudor M, Fambrough D, Csankovszki G, Jaenisch R
(2001) Loss of genomic methylation causes p53-dependent apoptosis and epigenetic deregula-
tion. Nat Genet 27:31–39
Jang HS, Shin WJ, Lee JE, Tae Do J (2017) CpG and non-CpG methylation in epigenetic gene
regulation and brain function. Genes (Basel) 8:148
Jefferson WN, Chevalier DM, Phelps JY, Cantor AM, Padilla-Banks E, Newbold RR, Archer TK,
Kinyamu HK, Williams CJ (2013) Persistently altered epigenetic marks in the mouse uterus
after neonatal estrogen exposure. Mol Endocrinol 27:1666–1677
Jones P, Baylin S (2007) The epigenomics of cancer. Cell 128:683–692
Josefowicz SZ, Wilson CB, Rudensky AY (2009) Cutting edge: TCR stimulation is sufficient for
induction of Foxp3 expression in the absence of DNA methyltransferase 1. J Immunol
182:6648–6652
Kaplan MJ, Beretta L, Yung RL, Richardson BC (2000) LFA-1 overexpression and T cell
autoreactivity: mechanisms. Immunol Investig 29:427–442
Kautiainen TL, Jones PA (1986) DNA methyltransferase levels in tumorigenic and nontumorigenic
cells in culture. J Biol Chem 261:1594–1598
Klose RJ, Bird AP (2006) Genomic DNA methylation: the mark and its mediators. Trends Biochem
Sci 31:89–97
Knox JD, Araujo FD, Bigey P, Slack AD, Price GB, Zannis-Hadjopoulos M, Szyf M (2000)
Inhibition of DNA methyltransferase inhibits DNA replication. J Biol Chem 275:17986–17990
Kohli RM, Zhang Y (2013) TET enzymes, TDG and the dynamics of DNA demethylation. Nature
502:472–479
Koike N, Yoo SH, Huang HC, Kumar V, Lee C, Kim TK, Takahashi JS (2012) Transcriptional
architecture and chromatin landscape of the core circadian clock in mammals. Science
338:349–354
Kurian JR, Terasawa E (2013) Epigenetic control of gonadotropin releasing hormone neurons.
Front Endocrinol 4:61
Lee PP, Fitzpatrick DR, Beard C, Jessup HK, Lehar S, Makar KW, Cherry SR (2001) A critical role
for Dnmt1 and DNA methylation in T cell development, function, and survival. Immunity
15:763–774
Lei H, Oh SP, Okano M, Juttermann R, Goss KA, Jaenisch R, Li E (1996) De novo DNA cytosine
methyltransferase activities in mouse embryonic stem cells. Development 122:3195–3205
312 C. S. Coyle et al.
Westwood FR (2008) The female rat reproductive cycle: a practical histological guide to staging.
Toxicol Pathol 36:375–384
Widschwendter M, Jones A, Evans I, Reisel D, Dillner J, Sundstrom K, Steyerberg EW,
Vergouwe Y, Wegwarth O, Rebitschek FG, Siebert U, Sroczynski G, de Beaufort ID, Bolt I,
Cibula D, Zikan M, Bjorge L, Colombo N, Harbeck N, Dudbridge F, Am T, Knoppers BM,
Joly Y, Teschendorff AE, Pashayan N, FORECEE 4C Consortium (2018) Epigenome-based
cancer risk prediction: rationale, opportunities and challenges. Nat Rev Clin Oncol 15:292–309
Wood GA, Fata JE, Watson KLM, Khokha R (2007) Circulating hormones and estrous stage
predict cellular and stromal remodelling in murine uterus. Reproduction 133:1035–1044
Xu HJ, Cairns P, Hu SX, Knowles MA, Benedict WF (1993) Loss of RB protein expression in
primary bladder cancer correlates with loss of heterozygosity at the RB locus and tumor
progression. Int J Cancer 53:781–784
Yeo GS, Heisler LK (2012) Unraveling the brain regulation of appetite: lessons from genetics.
Nat Neurosci 15:1343–1349
Yoder JA, Soman NS, Verdine GL, Bestor TH (1997) DNA (cytosine-T)-methyltransferase in
mouse cells and tissues. Studies with a mechanism-based probe. J Mol Biol 270:385–395
Part III
Development of Neuroendocrine Circuits
Development of Limbic System
Stress-Threat Circuitry 12
Newton S. Canteras, Dayu Lin, and Joshua G. Corbin
Abstract
How an animal responds to environmental stressors and threats is essential for
survival. These responses are governed through an interconnected circuit in the
brain dubbed the limbic system. Three main structures of the limbic system are
the amygdala, bed nucleus of the stria terminalis (BNST), and hypothalamus.
Together these structures define a central stress-threat circuit. This chapter
describes the most up to date knowledge of how these structures function in
regulating responses to stressors and how each of these structures is formed from
embryonic development. This knowledge of the underlying biology of these
regions is essential to design rational therapeutic approaches for conditions,
such as post-traumatic stress disorder (PTSD) that are characterized by
dysregulation of the stress-threat system. In this chapter, we also describe the
critical unanswered questions in the field of stress-threat research and potential
future research directions.
N. S. Canteras
Department of Anatomy, Institute of Biomedical Sciences, University of São Paulo; São Paulo, SP,
Butanta, Brazil
D. Lin
Neuroscience Institute, New York University School of Medicine, New York, NY, USA
Department of Psychiatry, New York University School of Medicine, New York, NY, USA
Center for Neural Science, New York University, New York, NY, USA
J. G. Corbin (*)
Center for Neuroscience Research, Department of Pediatrics, Children’s Research Institute,
Children’s National Medical Center, Washington, DC, USA
e-mail: [email protected]
Keywords
Stress-threat system · Amygdala · Bed nucleus of stria terminalis ·
Hypothalamus · Embryonic development · Brain circuitry · Defense responses ·
Loss and gain of function · Mouse models · Circuit manipulation
12.1 Introduction
Research into the neural mechanisms underlying defensive responses in animals has
a long history. Animals are naturally selected to protect themselves from threats
associated with the presence of a predator or aggressive conspecifics that evoke
innate fear. Evidence accumulated over the past two decades suggests that the
circuitry that supports fear responses is complex and involves multiple independent
circuits that process different types of fear. In particular, there is good evidence to
support the existence of distinct circuits for fear of predators and fear of aggressive
conspecifics (a member of the same species). In Sect. 12.2 of this chapter, we review
these data and consider these segregated fear circuits and point out the crucial role of
the amygdala, hypothalamus and bed nuclei of the stria terminalis. In Sect. 12.3 of
this chapter, we review the current state of knowledge of how these brain regions are
developmentally formed, and in Sect. 12.4 we describe the critical unanswered
questions in this field and future directions.
Fig. 12.1 Parallel circuits mediating innate responses to predators and aggressive conspecifics. See
text for details. Abbreviations: BMAp basomedial amygdalar nucleus, posterior part, BNSTif, pr bed
nuclei stria terminalis, interfascicular nucleus, principal nucleus; LA latereal amygdalar nucleus,
MEApv, pd medial amygdalar nucleus, posteroventral and posterodorsal parts, MPN medial
preoptic nucleus; PAGdm, dl, l periaqueductal gray, dorsomedial, dorsolateral and lateral parts,
PMDdm, vl dorsal premammillary nucleus, dorsomedial and ventrolateral parts, PMV ventral
preamammillary nucleus; VMHdm, vl ventromedial hypothalamic nucleus, dorsomedial and ven-
trolateral parts
(continued)
12 Development of Limbic System Stress-Threat Circuitry 321
(continued)
322 N. S. Canteras et al.
In the hypothalamus, it has been recognized that behavioral control columns, formed
by distinct hypothalamic circuits, are involved in the control of the three basic
classes of goal-oriented behaviors: ingestive, reproductive, and anti-predatory defen-
sive (Swanson 2000). This concept was mostly based on a systematic analysis of
axonal projections from the medial half of the hypothalamus of the rat and was
further corroborated by a number of functional studies (see Swanson 2000). Based
on anatomical tract-tracing experiments of medial hypothalamic zone projections,
this zone has been divided into two networks that show a high degree of intercon-
nection among nuclei (Fig. 12.1, and see Box 1 for description of modern tools now
being employed to understand the link between circuity and behavior). One network
comprises the anterior hypothalamic nucleus (AHN), the dorsomedial part of the
ventromedial nucleus (VMHdm), and the ventrolateral part of the dorsal
premammillary nucleus (PMDvl), which are highly interconnected (Canteras and
Swanson 1992; Canteras et al. 1994; Risold et al. 1994) (Fig. 12.1). Exposure to a
predator or its odor activates the AHN–VMHdm–PMDvl circuit (i.e., the predator-
responsive circuit; Fig. 12.1) (Cezario et al. 2008). Notably, the VMHdm and the
AHN provide a unique type of projection to the PMdvl, with dense bilateral
projection fields (Canteras et al. 1994; Risold et al. 1994). This type of synaptic
arrangement enables the PMdvl to work as an amplifier for the hypothalamic neural
processing of predator-related cues. Corroborating this view, the PMDvl is the most
responsive brain site during exposure to a cat or its odor, and lesions or pharmaco-
logical blockade of the PMD drastically reduces fear responses to a predator or its
odor (see Cezario et al. 2008). The VMHdm, in turn, receives most of the amygdalar
processing of predator-related cues from the MEApv (Canteras et al. 1995) and the
12 Development of Limbic System Stress-Threat Circuitry 323
BMAp (Petrovich et al. 1996) (Fig. 12.1). In this regard, of particular relevance is the
finding that exposure to different types of carnivore odors converge in a pathway
formed by the MEApv and VMHdm (Pérez Gómez et al. 2015). Pharmacogenetic
inhibition in transgenic mice expressing hM4D in steroidogenic factor 1 (SF1)-
expressing neurons in the central and dorsomedial parts of the VMH decreased the
defensive responses of the mice during exposure to a rat but not during exposure to
an aggressive conspecific (Silva et al. 2013). Conversely, optogenetic activation of
SF1-expressing neurons in the same population of VMH cells initiates a range of
behavioral responses that resemble the animal’s natural defensive behaviors (Wang
et al. 2015). The main targets of the VMHdm are the dorsal periaqueductal gray
(PAGd) and the AHN. Accordingly, activation of the pathway from the VMHdm to
the PAG induces immobility, whereas stimulation of the pathway from the VMHdm
to the AHN promotes mobility, running, escape jumping, and avoidance and the
animals tend to avoid the place where they had received the stimulation (Wang et al.
2015). At least part of the complex defensive behaviors that have been described for
the AHN are likely to be mediated through its projections to the PMDvl. The PMDvl
is the main interface between the predator-responsive circuit and the PAG (Motta et
al. 2009-Key Reference). The PMDvl provides strong projections to the dorsomedial
and dorsolateral parts of the PAG, both of which are activated during exposure to a
live predator or its odor and is an important site for the organization of fear responses
to predators (Motta et al. 2009-Key Reference) (Fig. 12.1). Notably, lesions in the
dorsal PAG block the entire range of fear responses to a predator, from flight and
freezing to risk assessment, and is conceivable that the degree of activation of the
dorsal PAG dictates the intensity of the fear response (see Cezario et al. 2008).
The second medial hypothalamic network includes the medial preoptic nucleus
(MPN), the ventrolateral part of the VMH (VMHvl), the tuberal nucleus, and the
ventral premammillary nucleus (PMV), which are also highly interconnected
(Simerly and Swanson 1988; Canteras et al. 1992, 1994). In contrast to predator
exposure, exposure to an aggressive conspecific activates the MPN–VMHvl–PMV
network (the conspecific-responsive circuit; Fig. 12.1), clearly suggesting the exis-
tence of separate circuits responsive to predators and conspecifics (see Motta et al.
2009-Key Reference). In addition, exposure to an aggressive conspecific activates
the dorsomedial part of the PMD (PMDdm), which is distinct from the predator-
activated PMDvl (Fig. 12.1). Of particular relevance, lesions of the PMD block
freezing and passive supine responses to an aggressive conspecific but not, however,
active defense, including boxing and upright defensive postures (Motta et al. 2009-
Key Reference). These data suggest that separate circuits in the PMD are responsible
for fear responses to predators versus fear responses to aggressive conspecifics. The
link between elements of the conspecific-responsive hypothalamic circuit and the
PMDdm is not clear at the moment, and the subfornical region of the lateral
hypothalamic area has been suggested one possible candidate (Goto et al. 2005).
As discussed for the predator-responsive medial hypothalamic circuit, the PMD
serves as an important interface between the conspecific-responsive hypothalamic
circuit and the PAG. However, differently from the PMDvl, the PMDdm projects to
the dorsomedial PAG (PAGdm) and lateral PAG (PAGl) (Fig. 12.1), providing a
324 N. S. Canteras et al.
Table 12.1 Summary of gain and loss of functional studies of the defense circuit
Area (cell type) Manipulation Behavioral change References
VMHdm (SF1+ ChR2 activation Freeze, flight, jump Wang et al. (2015)
cells) Kunwar et al. (2015)
VMHdm(SF1+ DREADDi Reduced defense toward a Silva et al. (2013)
cells) inactivation predator
VMHdm(SF1+ Capase3 mediated Reduced predator Kunwar et al. (2015)
cells) cell ablation avoidance
VMHdm ChR2 activation Freeze, flight, jump Wang et al. (2015)
(SF1)- > AHN
VMHdm ChR2 activation Freeze Wang et al. (2015)
(SF1)- > PAG
AHN ChR2 activation Flight and jump Wang et al. (2015)
(nonselective)
VMHvl DREADDi Reduced social defense Silva et al. (2013)
(nonselective) inactivation
VMHvl (Esr1+) ChR2 activation Increase social defense Wang et al. (2019)
VMHvl (Esr1+) NpHR3.0 Decrease active social Wang et al. (2019)
inactivation defense
PMd Cytotoxic lesion Decrease passive social Motta et al. (2009)-
(nonselective) defense Key Reference
PAGdm Cytotoxic lesion Decrease passive and active Canteras lab,
(nonselective) social defense unpublished results
projection pattern that matches the PAG activation pattern in response to an aggres-
sive conspecific (see Motta et al. 2009-Key Reference). Preliminary findings from
the Canteras laboratory suggest that PAGdm cytotoxic lesions decrease both passive
(freezing and the typical on-the-back position maintained after the resident leaves
them alone) and active forms (an upright position with sparse boxing and dashing
away from the resident) of social defensive responses. In addition to the PMD,
pharmacogenetic inhibition in the VMHvl of mice locally infected with an adeno-
associated virus (AAV) expressing hM4D pharmacogenetic neural inhibition tool
(HA-hM4D) decreased the defensive responses of the mice during exposure to an
aggressive conspecific but not during exposure to a rat (Silva et al. 2013). Moreover,
estrogen receptor+ (Esr1+) cells situated at the anterior portion of the VMHvl are
significantly excited when animals actively defended themselves against conspecific
aggressors. Furthermore, during exposure to a conspecific aggressor, inactivation of
VMHvlEsr1 cells compromised the active forms of social defensive behaviors, such
as dashing away from the aggressor and assuming an upright posture (Wang et al.
2019). A summary of functional studies implicating the VMH in the defensive
behaviors is shown in Table 12.1.
12 Development of Limbic System Stress-Threat Circuitry 325
The BNST serves as an important interface between the amygdala and the hypotha-
lamic circuits involved in social and anti-predatory defense (Fig. 12.1). However, the
role the BNST plays in the context of social and anti-predatory defense remains to be
fully investigated. Previous research has established the role of the BNST in
sustained fear as distinct from the role of the amygdala in phasic fear (see Lebow
and Chen 2016; Davis et al. 2010). Accordingly, the amygdala is likely to be
generically involved in the assessment of specific cues related to fearful events,
whereas the BNST would be related to the ability to differentiate between safe and
non-safe situations (Sullivan et al. 2004), causing a sustained state of apprehension
(see Lebow and Chen 2016; Davis et al. 2010). On hodological grounds, the
elements of the posterior division of the BNST fit quite well with the parallel circuits
organizing anti-predatory and social defense. Thus, in the posterior BNST, the
principal nucleus is particularly targeted by the MEApd (Canteras et al. 1995),
which, as previously mentioned, responds to aggressive conspecifics, and projects
to elements of the conspecific-responsive hypothalamic circuit the MPN, VMHvl,
tuberal nucleus, and PMV (Dong and Swanson 2004) (Fig. 12.1). Conversely, the
interfascicular nucleus of the posterior division of the BNST (BNSTif) receives
amygdalar inputs from regions processing predator cues, namely the MEApd and
BMAp (Canteras et al. 1995; Petrovich et al. 1996) and projects to elements of the
predator-responsive hypothalamic circuit, including the AHN, the VMHdm and, to a
lesser degree, the PMD (Dong and Swanson 2004) (Fig. 12.1). Moreover, the
interfascicular and principal nuclei of the BNST receive strong inputs from elements
conspecific- and predator-responsive medial hypothalamic circuits (Simerly and
Swanson 1988; Canteras et al. 1992, 1994; Risold et al. 1994), suggesting that the
BNST has more integrative functions that should be investigated in the context of
social and anti-predatory defense.
As described above many different brain structures and nuclei comprise the stress/
threat system. In this section we cover what is currently understood regarding
development of three central brain regions of this system: (1) the ventromedial
hypothalamic (VMH), paraventricular (PVN), and premammilary (PMN) nuclei of
the hypothalamus, (2) the MEA, and (3) the BNST. These brain structures are highly
tuned to detecting and processing environmental threats, and as such make up a core
interconnected circuit regulating stress/threat responses. While the knowledge of
development of each system remains incomplete and lags behind that of much more
studied brain regions such as the cerebral cortex, a picture is emerging of how each
structure is assembled during embryogenesis and how these developmental
programs may relate to how they function. Below we describe development of
each brain structure individually (Sect. 12.3.1–12.3.3), followed by what is known
about how connectivity is established (Sect. 12.3.4), and conclude with data and
326 N. S. Canteras et al.
Fig. 12.2 BNST and MEA progenitor zones. Schematics of the E13.5 brain at the level of the
telencephalon and diencephalon showing both the overall anatomy and regions of defined progeni-
tor zones known to contribute to either BNST or MEA (A & B). Table below in (C) lists the
transcription factors expressed in each progenitor zone and contributions to either the BNST or
MEA. Abbreviations: CGE caudal ganglionic eminence, CVP caudal-ventral pallium, Di Dienceph-
alon, dLGE dorsal lateral ganglionic eminence, DP dorsal pallium, LGE lateral ganglionic emi-
nence, LP lateral pallium, LV lateral ventricle, MGE medial ganglionic eminence, POA preoptic
area, VP ventral pallium, 3v third ventricle (Illustrations by Katie Sokolowski)
speculation regarding how these developmental genetic programs may link to later
stress/threat behavioral responses in adult animals (Sect. 12.3.5).
The major hypothalamic subnuclei that regulate threat avoidance and detection are
the VMH, the PVN, and the PMN (Gross and Canteras 2012-Key Reference;
Hashikawa et al. 2016-Key Reference; Takahashi 2014) (Fig. 12.1). As development
of these regions are covered in greater detail in the chapters by M. Placzek and M
Towers (Development of the Neuroendocrine Hypothalamus), and J. Michaud
(Development of the Paraventricular Nucleus), here we provide a general overview,
as readers should refer to the above chapters as well as outstanding current reviews
(Bedont et al. 2015; Alvarez-Bolado 2018; Xie and Dorsky 2017) for more details.
The vast majority of neurons in the hypothalamus are generated during embry-
onic development between embryonic day (E)10.5 and E16.5 in mouse (Shimada
12 Development of Limbic System Stress-Threat Circuitry 327
and Nakamura 1973 and reviewed in Bedont et al. 2015) (Mouse gestation is
approximately 19 days). Hypothalamic progenitor cells are born in the ventricular
zone surrounding the ventral aspect of the third ventricle (di on Fig. 12.2). Dividing
progenitors, spatially distributed along both dorsal-ventral and rostral-caudal axes of
the developing diencephalon, express different combinations of transcription factors
(Ferran et al. 2015; Shimogori et al. 2010), master regulator genes that function in a
combinatorial manner to determine cell fate. Differential transcription factor expres-
sion is one of the early developmental steps in which unspecified neural progenitors
progressively take on neuronal subtype identity prior to achieving their mature
neuronal phenotype. Table 12.2 summarizes the key transcription factors that com-
prise the unique “combinatorial code” for development and specification of major
subclasses of neurons of the VMH, PVN, and PMN (Bedont et al. 2015; Alvarez-
Bolado 2018; Xie and Dorsky 2017; Nouri and Awatramani 2017). These genes are
expressed during the proliferative and/or early postmitotic stages of hypothalamic
neuron development. It is important to note that Table 12.2 represents an incomplete
catalog of genes required for specification of neurons in these nuclei, especially the
PMN, as a number of required transcription factors remain to be identified.
At the anatomical level, the MEA can be subdivided into three distinct subdomains:
Anterior (MEAA), posterio-dorsal (MePd) and posterior-ventral (MePv). At the
neuronal subtype level, the MEA is composed of a mix of three major classes of
neurons: local projecting interneurons (Bian 2013; Guirado et al. 2008; Real et al.
2009), and long-range projecting excitatory and inhibitory output neurons (Bian et
al. 2008; Choi et al. 2005; Keshavarzi et al. 2014). Single-cell RNA-sequencing
studies have revealed that the MEA is populated by at least 16 distinct molecularly
identifiable neuronal subtypes (Wu et al. 2017). Much of this diversity is found
across interneurons, which based on a combination of previously identified molecu-
lar, electrophysiological and morphological criteria, fall into a number of distinct
subtypes (Petilla Interneuron Nomenclature Group et al. 2008). In the MEA, these
include subtypes that are defined by their expression of either somatostatin (Sst),
Calbindin (CB), Calretinin (CR), the serotonin receptor subunit 5Ht3a, or Neuro-
peptide Y (NPY). Interestingly, the MEA is devoid of interneurons marked by
328 N. S. Canteras et al.
expression of parvalbumin (PV) (Carney et al. 2010; Guirado et al. 2008), a subtype
that is found in abundance in the cerebral cortex.
In addition to these local and projection inhibitory neurons, the MEA is
comprised of excitatory output neurons. This mixing of inhibitory and excitatory
output neurons within the same structure is atypical for telencephalic nuclei as output
neurons of telencephalic nuclei are typically either excitatory (e.g., cerebral cortex,
basolateral amygdala) or inhibitory (e.g., basal ganglia), but not both. MEA neurons
can also be identified based on expression of one or more neuropeptides and
neurohormones, factors which mediate the internal state of the animal and appropri-
ate behavioral responses to stressors and threats (Gross and Canteras 2012-Key
Reference; Hashikawa et al. 2018; Li and Dulac 2018; Sokolowski and Corbin
2012; Yang and Shah 2014). How expression of these factors relates to the other
criteria for defining neuronal identity in the MEA remains an area of active investi-
gation. Although less well characterized than their interneuron counterparts, both
inhibitory and excitatory output neurons in the MEA can also be classified according
to distinct intrinsic electrophysiological profiles (Bian 2013; Bian et al. 2008;
Carney et al. 2010; Keshavarzi et al. 2014).
TRANSPLANTATION
E13.5 GFP Donor E13.5 non-GFP Recipient Adult Recipient
ELECTROPORATION
Plasmid E13.5 non-GFP Recipient Adult Recipient
E GFP
TRANSGENIC
F0 Generaon F1 Generaon
dLGE: Pax6
POA: Dbx1
CGE: Coup-TFII
B
MEA DP/LP: Emx1
CGE: Coup-TFII
CVP: Ebf3
Fig. 12.3 Adult BNST and MEA diversity. Coronal sections of the brain at the level of the BNST
(A, green) and MEA (B, blue) show relationship between gene expression patterns in embryonic
progenitor domains and neuronal diversity in each structure. See Fig. 12.1 for abbreviations
(Illustrations by Katie Sokolowski)
The above-described complexity of MEA neuronal cell types may be further under-
stood through the perspective of embryonic development when neurons are first
specified. A relatively significant amount of work studying dynamic gene expression
changes over developmental time, short-and long-term cell tracking experiments
using lipophilic fluorescent dyes, in vivo neuronal labeling, embryonic transplanta-
tion and genetic fate mapping has provided a relatively comprehensive map for
development of diverse MEA subpopulations (Fig. 12.3, Table 12.3, and references
therein, and see Box 2 for description of tracing techniques). From these studies, ~9
of the at least 16 MEA neuronal subclasses can be identified by differential devel-
opmental origin and/or embryonic transcription factor expression (Table 12.3, Figs.
12.2 and 12.3). At the broadest anatomical level, the three major MEA subnuclei
(MEAA, MePD, and MePV) can be individually identified in the postnatal brain by
the differential expression of members of the LIM family of homeodomain encoding
transcription factors, Lhx5, Lhx6, and Lhx9 (Choi et al. 2005-Key Reference; García-
López et al. 2008). The expression of these genes is atypical of developmentally
12 Development of Limbic System Stress-Threat Circuitry 331
factor encoding genes such as Tbr1, Foxp2, Nkx2.1, and Otp (Table 12.3, and
references therein). Previous fate mapping studies using Cre-based approaches
(see Box 2) have begun to uncover the relationship between neuronal progenitor
cell identity as defined by transcription factor expression and ultimate MEA neuro-
nal identity. While a full mapping of MEA neuronal diversity remains to be
accomplished and subsequently linked back to developmental origin, there are
clear patterns that have begun to emerge. First, fate mapping work, along with
previous standard birthdating studies, revealed that MEA neuronal specification
occurs within a similar epoch as the hypothalamus, roughly between E10.5 and
E15.5 in the mouse (Soma et al. 2009). Second, the same ventral telencephalic
domains (MGE and CGE), which generate interneurons of the cerebral cortex
(Batista-Brito and Fishell 2009; Xu et al. 2003), are also major sources of interneu-
ron populations for the MEA (Figs. 12.2, 12.3 and Table 12.3, and references
therein). For example, the CGE is the source of calretinin+ neurons and the MGE
is the source of Sst + and Calbindin+ interneurons in the MEA (Kanatani et al. 2015;
Nery et al. 2002; Soma et al. 2009; Touzot et al. 2016). Moreover, similar to the
cerebral cortex, a population of MEA excitatory neurons appears to be derived from
Emx1+ progenitors originating in lateral and dorsal pallial compartments of the
cerebral cortical primordium (Gorski et al. 2002) (Figs. 12.2, 12.3 and Table 12.3).
In contrast to this “shared” origin with other telencephalic brain regions for
populations of MEA interneurons and excitatory neurons, there appear to be multiple
unique sources of origin for MEA neuronal diversity. These include progenitor pools
located embryonic zones that straddle both sides of the telencephalic–diencephalic
border. The telencephalic niches include the preoptic area (POA) (Bulfone et al.
1993) (not be confused with the hypothalamic preoptic area, which by all evidence
does not appear to be derived from the embryonic POA), and likely a region known
as the entopeduncular region (AEP) (Fig. 12.2). These niches are located ventral to
the MGE and CGE, respectively, and are the ventral-most telencephalic domains.
Genetic fate mapping studies have revealed contributions of Nkx5.1+ POA-derived
neurons to the MEA (Although their adult identity was not established) (Gelman et
al. 2011). In addition, more in depth fate mapping studies have revealed Dbx1+ and
Foxp2+ contributions to the MEA from domains at the telencephalic–diencephalic
boundary. As defined by molecular and electrophysiological profiles, Dbx1- and
Foxp2-derived populations generate distinct subclasses of putative inhibitory output
neurons that are distributed across different subnuclei of the MEA (Lischinsky et al.
2017). These developmentally defined subpopulations also contribute to known
neuroendocrine-expressing populations such as those that express Aromatase,
Androgen Receptor (AR), and Estrogen receptor-α (ERα) (Table 12.3). Addition-
ally, as one moves caudally in the telencephalon there is a distinct compartment
where the pallium appears to wrap underneath the CGE, dubbed the caudo-ventral
pallium (CVP) (Ruiz-Reig et al. 2018) (Fig. 12.2). The CVP expresses the transcrip-
tion factor encoding gene, Ebf3, and the secreted factor encoding genes, Gdf10 and
Fgf15. These populations contribute to the MEA and appear to generate a subset of
excitatory neurons. This segregation of excitatory and inhibitory sources for MEA
neuronal diversity follows the current dogma that the subpallial telencephalon is the
12 Development of Limbic System Stress-Threat Circuitry 333
Located within the ventral telencephalon, the BNST, sometimes also referred to as
“the extended amygdala” is bounded by the globus pallidus and traverses from the
334 N. S. Canteras et al.
septal area to anterior to the hypothalamus (Figs. 12.2 and 12.3). It is composed of up
to 18 distinct subnuclei which are broadly grouped as either part of the anterior or
posterior BNST (Lebow and Chen 2016-Key Reference). As revealed largely from
the work of Larry Swanson and collaborators, the BNST sends and receives
projections from a variety of brain regions, most notably major subnuclei of the
amygdala and hypothalamus (Dong et al. 2001; Thompson and Swanson 1998). At
the neuronal level, the BNST expresses a large variety of neurohormones and
neuropeptides, which points to its central role in modulation of mood, anxiety, and
stress responses. Unfortunately, little work has been undertaken to understand how
this complex and important structure arises, thus BNST development remains a
critical understudied area. However, from a handful of studies, some insight has been
gained. First, classical Brdu birthdating studies from Shirley Bayer, a pioneer along
with Joseph Altman in the use of methodologies to birthdate neurons, showed that
cells that will populate the BNST are born across a rostral to caudal gradient (rostral
earlier) between E13 and E20 in the rat (Bayer 1987). Subsequent studies by George
De Vries have revealed distinct temporal windows within this period for the sexual
dimorphic development of vasopressin (Avp) + and Galanin (Gal+) cells during the
earlier waves of BNST generation (Han and De Vries 1999; al-Shamma and De
Vries 1996). Following Brdu labeling over time the work from Bayer (Bayer 1987)
suggested that BNST cells arise from the medial edge of the MGE near the septum,
and a more dorsal and posterior source located at the approximate level of the CGE
(Note while the term “ganglionic eminences” were first used in the 1970s, it was not
until the early 1990s that the terms MGE, LGE, and CGE were in use). Following
these early studies, more recent gene expression studies carried out primarily from
the laboratories of Loreta Medina and Luis Puelles have extended our understanding
of BNST development (Bupesh et al. 2011; García-López et al. 2008; Puelles et al.
2000). This work, combined with work of others (Hirata et al. 2009; Nery et al. 2002;
Ruiz-Reig et al. 2018; Touzot et al. 2016; Vucurovic et al. 2010), indicates that
patterning of the BNST follows similar rules as MEA development with multiple
sources contributing to the BNST (Table 12.5, and references therein). These
sources likely include the CGE, POA, MGE (AEP), LGE, and perhaps the ventral
pallium (VP) (Figs. 12.2, 12.3 and Table 12.5). However, considering the wide
variety of neuronal cell types that populate the BNST, there are likely other sources
including the dorsal telencephalon as well as extratelencephalic sources (Bupesh et
al. 2011).
12 Development of Limbic System Stress-Threat Circuitry 335
The above subsections detail known mechanisms of how diverse neurons of key
threat detection structures are specified during embryonic development. However,
how these circuits become wired together still remains unclear. A number of lines of
evidence suggest that circuitry within this system is laid down relatively early. For
example, studies from the lab of Richard Simerly have shown that circuitry between
the BNST and MEA is already established by postnatal day (P)5 (Hutton et al. 1998).
Moreover, although not experimental evidence, reporter expression studies of limbic
system expressed genes reveal that neuronal tracts of the limbic system can be
observed as early as E15.5 (Fig. 12.4).
Fig. 12.4 Gensat images. DAB immunostained images of GFP expression from BAC driven cre
recombination revealing gene expression of the cell adhesion encoding genes Pchd20, Amigo2 or
Pchd9 by E15.5. Recombination is observed in the olfactory bulb (OB, with asterisk highlighting
AOB), amygdala and hypothalamus, with presumptive projections observed between the OB and
amygdala/hypothalamus (arrow). These images suggest establishment of limbic circuits as by as
early as E15.5. Images courtesy of the Gene Expression Nervous System Atlas (GENSAT) Project,
NINDS Contracts N01NS02331 & HHSN271200723701C to The Rockefeller University
(New York, NY). Abbreviations: Amy amygdala, AOB accessory olfactory bulb, Hyp hypothala-
mus, OB olfactory bulb.
The above subsections describe current (early 2019) knowledge of the circuity and
function of the stress/threat system (Sects. 12.2 and 12.3) and how neurons that
comprise key stress/threat systems are specified from development, with specific
examples of how key transcription factors establish subsystems required for select
behaviors (Sect. 12.4). Despite this knowledge, there remain a number of key open
questions, highlighted in Box 3 below:
Box 3 This Box Contains Interesting Information about Key Questions for
Future Research
Question #1: Do embryonic developmental programs contribute to sex
differences in stress/threat system behaviors, or is this solely governed by
later hormonal effects?
Importance: Many stress linked disorders such as PTSD and anxiety
disorders, as well as behavioral disorders characterized by altered stress
responses, such as autism spectrum disorders, display a sex bias in presenta-
tion and prevalence. In addition to addressing basic biological questions of
how male and female circuitry is established in sexually dimorphic brain
regions such as the BNST, MEA and hypothalamus, answering Question #1
will likely inform and possibly provide better therapeutic approaches to
treating these prevalent conditions.
Question #2: How do environmental stressors affect formation of these
“hardwired” systems?
Importance: While formation of BNST-MEA-Hypothalamic circuits
appears to be laid down primarily via developmental genetic programming
mechanisms, these circuits are highly malleable by environmental influences
during both fetal (embryonic) and early post-natal critical periods. Under-
standing the interplay between genes and environment is a critical unanswered
question that has implications for a variety of human conditions described
above in Question #1.
Question #3: Are there evolutionarily conserved common “blueprints” for
establishing stress/threat system connectivity?
Importance: Normal responses to immediate environmental threats allow
for survival of individuals across all species. This commonality may suggest
that there are evolutionarily conserved mechanisms for development of threat
systems across a diverse species, perhaps via utilization of similar genetic
modules for neuron and circuit specification. If so, this may expand the
repertoire of animal model systems in which to model human disorders
affecting stress/threat systems.
338 N. S. Canteras et al.
Key References
Choi et al. (2005)—This article identifies subcircuits of the medial amygdala that are
genetically identifiable by expression of the LIM-family of transcription factors
and posits that these subcircuits control different innate behaviors.
Gross and Canteras (2012)—This review nicely describes the brain circuits
controlling innate and learned fear.
Hirata et al. (2009)—Using genetic tracing techniques this study revealed that there
are progenitor zones in the developing telencephalon that are dedicated to
generate neuronal diversity in the amygdala.
Hashikawa et al. (2016)—This review eloquently describes the olfactory to deeper
brain circuits that translate odor information to innate behavioral output.
Lebow and Chen (2016)—This review provides an excellent summary the anatomy,
function and role of the bed nucleus of the stria terminalis in psychiatric disorders.
Luo, et al. (2018)—This review provides an very comprehensive description of the
state of the art tools of modern neuroscience.
Motta et al. (2009)—Using a combination of approaches, this study revealed that
there are distinct hypothalamic circuits for processing different fear behaviors.
References
al-Shamma HA, De Vries GJ (1996) Neurogenesis of the sexually dimorphic vasopressin cells of
the bed nucleus of the stria terminalis and amygdala of rats. J Neurobiol 29:91–98
Alvarez-Bolado G (2018) Development of neuroendocrine neurons in the mammalian hypothala-
mus. Cell Tissue Res 375(1):23–39
Amir-Zilberstein L, Blechman J, Sztainberg Y, Norton WHJ, Reuveny A, Borodovsky N, Tahor M,
Bonkowsky JL, Bally-Cuif L, Chen A et al (2012) Homeodomain protein otp and activity-
dependent splicing modulate neuronal adaptation to stress. Neuron 73:279–291
Aravanis AM, Wang LP, Zhang F, Meltzer LA, Mogri MZ, Schneider MB, Deisseroth K (2007) An
optical neural interface: in vivo control of rodent motor cortex with integrated fiberoptic and
optogenetic technology. J Neural Eng 4:S143–S156
Armbruster B, Roth B (2005) Creation of designer biogenic amine receptors via directed molecular
evolution. Neuropsychopharmacology 30:S265–S265
Armbruster BN, Li X, Pausch MH, Herlitze S, Roth BL (2007) Evolving the lock to fit the key to
create a family of G protein-coupled receptors potently activated by an inert ligand. Proc Natl
Acad Sci U S A 104:5163–5168
Batista-Brito R, Fishell G (2009) The developmental integration of cortical interneurons into a
functional network. Curr Top Dev Biol 87:81–118
Bayer SA (1987) Neurogenetic and morphogenetic heterogeneity in the bed nucleus of the stria
terminalis. J Comp Neurol 265:47–64
Bedont JL, Newman EA, Blackshaw S (2015) Patterning, specification, and differentiation in the
developing hypothalamus. Wiley Interdiscip Rev Dev Biol 4:445–468
Berndt A, Lee SY, Ramakrishnan C, Deisseroth K (2014) Structure-guided transformation of
Channelrhodopsin into a light-activated Chloride Channel. Science 3:420–424
Bian X (2013) Physiological and morphological characterization of GABAergic neurons in the
medial amygdala. Brain Res 1509:8–19
Bian X, Yanagawa Y, Chen WR, Luo M (2008) Cortical-like functional organization of the
pheromone-processing circuits in the medial amygdala. J Neurophysiol 99:77–86
12 Development of Limbic System Stress-Threat Circuitry 339
Ericson J, Muhr J, Jessell TM, Edlund T (1995) Sonic hedgehog: a common signal for ventral
patterning along the rostrocaudal axis of the neural tube. Int J Dev Biol 39:809–816
Fenno L, Yizhar O, Deisseroth K (2011) The development and application of optogenetics. Annu
Rev Neurosci 34:389–412
Ferran JL, Puelles L, Rubenstein JLR (2015) Molecular codes defining rostrocaudal domains in the
embryonic mouse hypothalamus. Front Neuroanat 9:46
Ferrero DM, Lemon JK, Fluegge D, Pashkovski SL, Korzan WJ, Datta SR et al (2011) Detection
and avoidance of a carnivore odor by prey. Proc Natl Acad Sci U S A 108:11235–11240
García-López M, Abellán A, Legaz I, Rubenstein JLR, Puelles L, Medina L (2008) Histogenetic
compartments of the mouse centromedial and extended amygdala based on gene expression
patterns during development. J Comp Neurol 506:46–74
García-Moreno F, Pedraza M, Di Giovannantonio LG, Di Salvio M, López-Mascaraque L, Simeone
A, De Carlos JA (2010) A neuronal migratory pathway crossing from diencephalon to telen-
cephalon populates amygdala nuclei. Nat Neurosci 13:680–689
Gelman D, Griveau A, Dehorter N, Teissier A, Varela C, Pla R, Pierani A, Marín O (2011) A wide
diversity of cortical GABAergic interneurons derives from the embryonic preoptic area. J
Neurosci 31:16570–16580
Gomez JL, Bonaventura J, Lesniak W, Mathews WB, Sysa-Shah P, Rodriguez LA, Ellis RJ, Richie
CT, Harvey BK, Dannals RF, Pomper MG, Bonci A, Michaelides M (2017) Chemogenetics
revealed: DREADD occupancy and activation via converted clozapine. Science 357:503–507
Gorski JA, Talley T, Qiu M, Puelles L, Rubenstein JLR, Jones KR (2002) Cortical excitatory
neurons and glia, but not GABAergic neurons, are produced in the Emx1-expressing lineage. J
Neurosci 22:6309–6314
Goto M, Canteras NS, Burns G, Swanson LW (2005) Projections from the subfornical region of the
lateral hypothalamic area. J Comp Neurol 493:412–438
Gradinaru V, Zhang F, Ramakrishnan C, Mattis J, Prakash R, Diester I, Goshen I, Thompson KR,
Deisseroth K (2010) Molecular and cellular approaches for diversifying and extending
optogenetics. Cell 141:154–165
Gross CT, Canteras NS (2012) The many paths to fear. Nat Rev Neurosci 13:651–658
Guirado S, Real MA, Dávila JC (2008) Distinct immunohistochemically defined areas in the medial
amygdala in the developing and adult mouse. Brain Res Bull 75:214–217
Han TM, De Vries GJ (1999) Neurogenesis of galanin cells in the bed nucleus of the stria terminalis
and centromedial amygdala in rats: a model for sexual differentiation of neuronal phenotype. J
Neurobiol 38:491–498
Hashikawa K, Hashikawa Y, Falkner A, Lin D (2016) The neural circuits of mating and fighting in
male mice. Curr Opin Neurobiol 38:27–37
Hashikawa K, Hashikawa Y, Lischinsky J, Lin D (2018) The neural mechanisms of sexually
dimorphic aggressive behaviors. Trends Genet TIG 34:755–776
Hegemann P, Nagel G (2013) From Channelrhodopsins to Optogenetics. EMBO Mol Med
5:173–176
Hirata T, Li P, Lanuza GM, Cocas LA, Huntsman MM, Corbin JG (2009) Identification of distinct
telencephalic progenitor pools for neuronal diversity in the amygdala. Nat Neurosci 12:141–149
Hutton LA, Gu G, Simerly RB (1998) Development of a sexually dimorphic projection from the
bed nuclei of the stria terminalis to the anteroventral periventricular nucleus in the rat. J
Neurosci 18:3003–3013
Keshavarzi S, Sullivan RKP, Ianno DJ, Sah P (2014) Functional properties and projections of
neurons in the medial amygdala. J Neurosci 34:8699–8715
Kunwar PS, Zelikowsky M, Remedios R, Cai H, Yilmaz M, Meister M, Anderson DJ (2015)
Ventromedial hypothalamic neurons control a defensive emotion state. elife 4:e06633
Kanatani S, Honda T, Aramaki M, Hayashi K, Kubo K, Ishida M, Tanaka DH, Kawauchi T, Sekine
K, Kusuzawa S, Kawasaki T, Hirata T, Tabata H, Uhlén P, Nakajima K (2015) The COUP-TFII/
Neuropilin-2 is a molecular switch steering diencephalon-derived GABAergic neurons in the
developing mouse brain. Proc Natl Acad Sci U S A 112(36):E4985–E4994. https://doi.org/10.
1073/pnas.1420701112. Epub 2015 Aug 24. PubMed PMID: 26305926; PubMed Central
PMCID: PMC4568674
12 Development of Limbic System Stress-Threat Circuitry 341
Kuerbitz J, Arnett M, Ehrman S, Williams MT, Vorhees CV, Fisher SE, Garratt AN, Muglia LJ,
Waclaw RR, Campbell K (2018) Loss of Intercalated Cells (ITCs) in the mouse amygdala of
Tshz1 mutants correlates with fear, depression, and social interaction phenotypes. J Neurosci 38
(5):1160–1177. https://doi.org/10.1523/JNEUROSCI.1412-17.2017. Epub 2017 Dec 18.
PubMed PMID: 29255003; PubMed Central PMCID: PMC5792476
Lebow MA, Chen A (2016) Overshadowed by the amygdala: the bed nucleus of the stria terminalis
emerges as key to psychiatric disorders. Mol Psychiatry 21:450–463
Li Y, Dulac C (2018) Neural coding of sex-specific social information in the mouse brain. Curr
Opin Neurobiol 53:120–130
Liberles SD (2015) Trace amine-associated receptors: ligands, neural circuits, and behaviors. Curr
Opin Neurobiol 34C:1–7
Lischinsky JE, Sokolowski K, Li P, Esumi S, Kamal Y, Goodrich M, Oboti L, Hammond TR,
Krishnamoorthy M, Feldman D et al (2017) Embryonic transcription factor expression in mice
predicts medial amygdala neuronal identity and sex-specific responses to innate behavioral cues.
elife 6:e21012
Luo L, Callaway EM, Svoboda K (2008) Genetic dissection of neural circuits. Neuron 57
(5):634–660
Luo LQ, Callaway EM, Svoboda K (2018) Genetic dissection of neural circuits: a decade of
Progress. Neuron 98:256–281
Madisen L, Mao T, Koch H, Zhuo JM, Berenyi A, Fujisawa S, Hsu YW, Garcia AJ 3rd, Gu X,
Zanella S, Kidney J, Gu H, Mao Y, Hooks BM, Boyden ES, Buzsaki G, Ramirez JM, Jones AR,
Svoboda K, Han X, Turner EE, Zeng H (2012) A toolbox of Cre-dependent optogenetic
transgenic mice for light-induced activation and silencing. Nat Neurosci 15:793–802
Magnus CJ, Lee PH, Bonaventura J, Zemla R, Gomez JL, Ramirez MH, Hu X et al (2019)
Ultrapotent Chemogenetics for research and potential clinical applications. Science 364
(6436):eaav5282. https://doi.org/10.1126/science.aav5282. [Epub ahead of print]
Martinez RC, Carvalho-Netto EF, Ribeiro-Barbosa ER, Baldo MV, Canteras NS (2011) Amygdalar
roles during exposure to a live predator and to a predator-associated context. Neuroscience
172:314–328
Matise MP, Wang H (2011) Sonic hedgehog signaling in the developing CNS where it has been and
where it is going. Curr Top Dev Biol 97:75–117
McGregor IS, Hargreaves GA, Apfelbach R, Hunt GE (2004) Neural correlates of cat odor-induced
anxiety in rats: region-specific effects of the benzodiazepine midazolam. J Neurosci
24:4134–4144
Motta SC, Goto M, Gouveia FV, Baldo MV, Canteras NS, Swanson LW (2009) Dissecting the
brain's fear system reveals the hypothalamus is critical for responding in subordinate conspecific
intruders. Proc Natl Acad Sci U S A 106:4870–4875
Nery S, Fishell G, Corbin JG (2002) The caudal ganglionic eminence is a source of distinct cortical
and subcortical cell populations. Nat Neurosci 5:1279–1287
Neve RL, Neve KA, Nestler EJ, Carlezon WA (2005) Use of herpes virus amplicon vectors to study
brain disorders. BioTechniques 39:381–391
Nouri N, Awatramani R (2017) A novel floor plate boundary defined by adjacent En1 and Dbx1
microdomains distinguishes midbrain dopamine and hypothalamic neurons. Dev Camb Engl
144:916–927
Pérez-Gómez A, Bleymehl K, Stein B, Pyrski M, Birnbaumer L, Munger SD et al (2015) Innate
predator odor aversion driven by parallel olfactory subsystems that converge in the ventromedial
hypothalamus. Curr Biol 25:1340–1346
Petilla Interneuron Nomenclature Group, Ascoli GA, Alonso-Nanclares L, Anderson SA,
Barrionuevo G, Benavides-Piccione R, Burkhalter A, Buzsáki G, Cauli B, Defelipe J et al
(2008) Petilla terminology: nomenclature of features of GABAergic interneurons of the cerebral
cortex. Nat Rev Neurosci 9:557–568
Petrovich GD, Risold PY, Swanson LW (1996) Organization of the projections of the basomedial
nucleus of the amygdala: a PHAL study in the rat. J Comp Neurol 374:387–420
342 N. S. Canteras et al.
Tervo DGR, Hwang BY, Viswanathan S, Gaj T, Lavzin M, Ritola KD, Lindo S, Michael S,
Kuleshova E, Ojala D, Huang CC, Gerfen CR, Schiller J, Dudman JT, Hantman AW, Looger
LL, Schaffer DV, Karpova AY (2016) A designer AAV variant permits efficient retrograde
access to projection neurons. Neuron 92(2):372–382
Thompson RH, Swanson LW (1998) Organization of inputs to the dorsomedial nucleus of the
hypothalamus: a reexamination with Fluorogold and PHAL in the rat. Brain Res Brain Res Rev
27:89–118
Touzot A, Ruiz-Reig N, Vitalis T, Studer M (2016) Molecular control of two novel migratory paths
for CGE-derived interneurons in the developing mouse brain. Dev. Camb. Engl.
143:1753–1765
Tye KM, Deisseroth K (2012) Optogenetic investigation of neural circuits underlying brain disease
in animal models. Nat Rev Neurosci 13:251–266
Tang K, Rubenstein JL, Tsai SY, Tsai MJ (2012) COUP-TFII controls amygdala patterning by
regulating neuropilin expression. Development 139(9):1630–1639. https://doi.org/10.1242/dev.
075564. PubMed PMID: 22492355; PubMed Central PMCID: PMC3317968
Vardy E, Robinson JE, Li C, Olsen RHJ, DiBerto JF, Giguere PM, Sassano FM, Huang XP, Zhu H,
Urban DJ, White KL, Rittiner JE, Crowley NA, Pleil KE, Mazzone CM, Mosier PD, Song J,
Kash TL, Malanga CJ, Krashes MJ, Roth BL (2015) A new DREADD facilitates the
multiplexed Chemogenetic interrogation of behavior. Neuron 86:936–946
Vucurovic K, Gallopin T, Ferezou I, Rancillac A, Chameau P, van Hooft JA, Geoffroy H, Monyer
H, Rossier J, Vitalis T (2010) Serotonin 3A receptor subtype as an early and protracted marker
of cortical interneuron subpopulations. Cereb Cortex N Y N 1991(20):2333–2347
Wang L, Chen IZ, Lin D (2015) Collateral pathways from the ventromedial hypothalamus mediate
defensive behaviors. Neuron 85:1344–1358
Wang L, Talwar V, Osakada T, Kuang A, Guo Z, Yamaguchi T, Lin D (2019) Hypothalamic
control of conspecific self-defense. Cell Rep 26:1747–1758.e5
Wircer E, Blechman J, Borodovsky N, Tsoory M, Nunes AR, Oliveira RF, Levkowitz G (2017)
Homeodomain protein Otp affects developmental neuropeptide switching in oxytocin neurons
associated with a long-term effect on social behavior. elife 6:e22170
Wu YE, Pan L, Zuo Y, Li X, Hong W (2017) Detecting activated cell populations using single-cell
RNA-Seq. Neuron 96:313–329.e6
Wyatt TD (2014) Pheromones and animal behavior, 2nd edn. Cambridge University Press,
Cambridge, UK
Xie Y, Dorsky RI (2017) Development of the hypothalamus: conservation, modification and
innovation. Dev Camb Engl 144:1588–1599
Xu Q, de la Cruz E, Anderson SA (2003) Cortical interneuron fate determination: diverse sources
for distinct subtypes? Cereb Cortex N Y N 1991(13):670–676
Yang CF, Shah NM (2014) Representing sex in the brain, one module at a time. Neuron
82:261–278
Zhao T, Szabó N, Ma J, Luo L, Zhou X, Alvarez-Bolado G (2008) Genetic mapping of Foxb1-cell
lineage shows migration from caudal diencephalon to telencephalon and lateral hypothalamus.
Eur J Neurosci 28:1941–1955
Zhu H, Aryal DK, Olsen RH, Urban DJ, Swearingen A, Forbes S, Roth BL, Hochgeschwender U
(2016) Cre-dependent DREADD (designer receptors exclusively activated by designer drugs)
mice. Genesis 54:439–446
Organization and Postnatal Development
of Visceral Sensory Inputs 13
to the Neuroendocrine Hypothalamus
Linda Rinaman
Abstract
Sensory signals arising in the body’s internal organs are delivered to the central
nervous system by spinal and vagal afferent neurons. These signals modulate
pituitary hormone release via ascending neural pathways from the caudal
brainstem to the neuroendocrine hypothalamus. Interoceptive (i.e., visceral) feed-
back regarding the moment-to-moment state of gastrointestinal and other physio-
logical systems is critical for shaping and coordinating adaptive neuroendocrine
functions, including hormonal responses to real or perceived homeostatic threats.
This chapter summarizes the functional organization and postnatal development
of visceral sensory inputs to the neuroendocrine hypothalamus in rats and mice,
the two mammalian models in which most of the relevant data have been collected.
A special emphasis is placed on hypothalamic inputs arising from noradrenergic
(NA) and glucagon-like peptide 1 (GLP1) neurons, whose cell bodies occupy the
caudal nucleus of the solitary tract (NTS) and medullary reticular formation (RF).
The axonal projections of NA and GLP1 neurons directly target the hypothalamic
medial preoptic area (mPOA), arcuate nucleus (ARC), paraventricular nucleus
(PVN), and supraoptic nucleus (SON), where adrenergic and GLP1 receptor
signaling modulates neuroendocrine gene expression, pulsatile activity, and pitui-
tary hormone release. NA and GLP1 inputs to the neuroendocrine hypothalamus
mature postnatally in rats and mice, likely contributing to the developmental
maturation of pituitary hormone release in response to both interoceptive and
exteroceptive signals.
Keywords
Fos · Gastrointestinal · Glucagon-like peptide 1 · HPA axis · Noradrenergic ·
Nucleus of the solitary tract · Oxytocin
L. Rinaman (*)
Department of Psychology and Program in Neuroscience, Florida State University, Tallahassee, FL,
USA
e-mail: [email protected]
13.1 Introduction
Sensory signals arising within the body’s internal organs (i.e., thoracic, abdominal,
and pelvic viscera) are carried to the CNS by spinal and cranial nerve afferents. In
rats, the cell bodies of spinal visceral sensory afferents occupy the dorsal root ganglia
13 Organization and Postnatal Development of Visceral Sensory Inputs to the. . . 347
Fig. 13.1 Mechanical and chemical signals arising within the stomach, intestines, and other
viscera are relayed to the caudal brainstem by vagal sensory afferents whose cell bodies occupy
the nodose ganglia (NG). The central axons of vagal afferents synapse within the nucleus of the
solitary tract (NTS) in the caudal medulla. Noradrenergic neurons (i.e., immunoreactive for
dopamine beta hydroxylase, DBH) and peptidergic neurons expressing preproglucagon and immu-
noreactive for glucagon-like peptide 1 (GLP1) are located within the NTS, reticular formation (RF),
and ventrolateral medulla (VLM), and carry visceral sensory signals from the caudal brainstem to
the neuroendocrine hypothalamus via the ventral noradrenergic bundle (VNAB). Lower left:
Postnatal changes in innervation of the neuroendocrine hypothalamus. In newborn rats and mice,
DBH- and GLP1-positive axons are sparse within the paraventricular nucleus (PVN) and other
hypothalamic endocrine nuclei, with innervation density increasing progressively to achieve adult-
like levels by the end of the third week postnatal (pink arrow)
(DRG); their axons terminate centrally in laminae I–VII of the spinal cord dorsal
horn and intermediate zone, and also in lamina X around the central canal. A subset
of the second-order spinal sensory neurons that receive visceral sensory inputs also
receive input from somatic sensory DRG neurons; thus, the activity of spinal sensory
neurons is co-modulated by visceral, cutaneous, and proprioceptive stimuli.
348 L. Rinaman
Given the apparent lack of direct spinal inputs to the neuroendocrine hypothalamus,
the balance of this chapter will focus on the structure and development of direct
hypothalamic inputs arising from the NTS and medullary RF (Fig. 13.1). In addition
to spinal sensory signals that are relayed via the spinosolitary tract, NTS neurons
receive direct visceral sensory input via two cranial nerves, the glossopharyngeal (IX)
and the vagus (X). The cell bodies of glossopharyngeal and vagal afferent neurons
occupy the inferior glossopharyngeal (aka petrosal) and nodose ganglia, respectively.
The peripheral axons of glossopharyngeal and a subset of vagal afferents monitor
cardiovascular and pulmonary target tissues within the thoracic cavity, while separate
populations of vagal afferents monitor the GI tract and associated digestive viscera,
from the oral cavity through the colon. The centrally projecting axons of
glossopharyngeal and vagal neurons enter the caudal dorsolateral medulla via multi-
ple sensory rootlets to converge within the solitary tract, which conveys visceral
sensory axons along the rostrocaudal length of the NTS. Most of these sensory
afferents form synapses within the caudal visceral portion of the NTS (as opposed
to the more rostral gustatory NTS), but some visceral sensory axons terminate within
the area postrema (AP), a dorsal midline circumventricular organ that lacks a blood-
brain barrier. The AP lies in the floor of the caudal fourth ventricle, immediately
adjacent to the visceral portion of the NTS. AP neurons respond to interoceptive
signals conveyed by hormones, osmolytes, toxins, cytokines, and other circulating
factors and convey these signals to postsynaptic neurons in the NTS, medullary RF,
and pontine parabrachial nucleus (PBN) (Shapiro and Miselis 1985).
Vagal sensory neurons within the rat nodose ganglia undergo their final mitotic
division between embryonic day (E)10 and E14, with peak cell production on E13.
Even as neural proliferation continues, the peripheral axons of vagal sensory neurons
start to innervate the GI tract, and their central axons begin to penetrate the caudal
13 Organization and Postnatal Development of Visceral Sensory Inputs to the. . . 349
Fig. 13.2 DiI tracing of vagal neurons during embryonic development in rats. DiI crystals were
placed into the developing gastric tube of formalin-fixed embryos harvested on embryonic day (E)
12 or E15. In the left panel, DiI-labeled vagal sensory neurons within the nodose ganglion (NG;
long arrows) are evident, along with their distal axons (asterisk) that are labeled due to their
peripheral association with the gastric tube. However, very few labeled NG neurons have axons
extending centrally into the caudal medulla (short arrows). In the right panel, DiI-labeled motor
neurons within the dorsal motor nucleus of the vagus (DMV) that innervate the gastric tube are
evident, along with the central axons of vagal sensory neurons that bundle within the solitary tract
(tr). Some of these sensory axons emerge to project into the developing NTS (nucleus tractus
solitarii, nucleus of the solitary tract; long arrows). Figure adapted from Rinaman and Levitt (1993)
medulla (Rinaman and Levitt 1993). Results from a study in which crystals of DiI
(1,10 -dioctadecyl-3,3,30 ,30 -tetramethylindocarbocyanine perchlorate) were placed
into the developing digestive tube in intact, formalin-fixed rat embryos (see Box
13.1) revealed that vagal sensory neurons within the nodose ganglia were first
labeled in E12 specimens (Fig. 13.2, NG), evidence that their peripheral axons
first reach the digestive tube at this time point. However, very few of the nodose
neurons labeled at E12 had central axons that entered the caudal medulla, and
labeled fibers within the solitary tract were not identifiable until E13. The density
and rostrocaudal distribution of labeled vagal afferents within the solitary tract
increased markedly between E13 and E14, but labeled vagal afferent axons did not
emerge from the tract until E15, when labeled fibers at all rostrocaudal levels coursed
directly from the solitary tract (Fig. 13.2, tr) into the proliferative ventricular zone of
the alar plate (Fig. 13.2, right, long arrows). Although vagal sensory fibers were not
observed to arborize within the developing NTS until E16, their distribution within
the NTS was remarkably adult-like by E18, approximately 3 days before birth
(Rinaman and Levitt 1993). Thus, the early embryonic development of vagal
sensory inputs to the NTS provides an anatomical substrate for digestive vago-
vagal reflexes (e.g., the gastric accommodation reflex) that already are operational
in newborn rats.
350 L. Rinaman
The ability of visceral sensory stimuli to modulate pituitary hormone release in adult
animals has been appreciated for many years. Early studies demonstrated posterior
pituitary release of AVP and OT in rats after electrical stimulation of the central end
of the severed vagus nerve (Chang et al. 1939; Thieblot et al. 1957), and later work
demonstrated that gastrointestinal or vagal nerve stimulation increases the firing
activity of magnocellular and parvocellular endocrine neurons within the PVN and
SON in anesthetized rats (Jin et al. 1993; Ueta et al. 1993). Some of these
experiments involved mechanical gastric distension or electrical stimulation of
vagal afferents, while others involved pharmacological manipulation of vagal
afferents via systemic administration of CCK or other gut peptides in adult rats.
For example, systemically administered ghrelin (a hormone that is endogenously
secreted from gastric tissue when the stomach is empty) suppresses the firing activity
of GI vagal afferents, whereas CCK increases vagal afferent firing. In adult rodents,
increased gastric release of ghrelin promotes increased release of growth hormone
(GH) from the anterior pituitary. Ghrelin binds to receptors expressed by vagal
sensory neurons, which transport the receptor to their peripheral axon terminals
within the stomach wall (see Fig. 13.1). Blockade of vagal afferent signaling to the
352 L. Rinaman
NTS abolishes ghrelin-induced GH secretion and also blocks the ability of ghrelin to
activate parvocellular GHRH endocrine neurons (Date et al. 2002).
Visceral sensory signals arising within the GI tract and other peripheral tissues are
relayed to the neuroendocrine hypothalamus via direct and indirect ascending neural
projection pathways from the NTS and AP in adult rats and mice (Rinaman 2007).
The indirect pathways include projections from AP and NTS neurons that are
relayed through the medullary RF (including the VLM) and pontine PBN before
ascending to the hypothalamus. Projections from AP neurons to the subjacent NTS
and also directly to the VLM and PBN provide a neural route for blood-borne signals
to alter central visceral sensory processing. However, while endocrine neurons
within the hypothalamic PVN, SON, mPOA, and ARC receive direct synaptic
input from catecholaminergic and peptidergic NTS and VLM neurons in adult
rats, they appear to receive little or no direct synaptic input from the pontine PBN
(Saper and Loewy 1980). The PBN instead appears to target primarily pre-auto-
nomic hypothalamic neurons and also provides robust input to the central nucleus of
the amygdala (CeA) and bed nucleus of the stria terminalis (BNST), two highly
interconnected limbic forebrain regions which also receive direct input from the NTS
and VLM (Bienkowski and Rinaman 2013).
2001). By P6, PrRP-positive fibers also are present within the PVN and other
hypothalamic nuclei, but fiber density is markedly sparser at P6 than in adulthood.
It is important to note that an absence or scarcity of neurotransmitter-positive
fibers should not be interpreted as evidence that the axons of NTS and VLM neurons
are not present within the hypothalamus. In one study in which retrograde neural
tracer was microinjected into the PVN in rat pups on the day of birth (P0), the
number of PVN-projecting neurons within the NTS and VLM quantified 24 h later
(on P1) was similar to the number of retrogradely labeled, PVN-projecting neurons
in adult rats (Rinaman 2001). However, a smaller proportion of PVN-projecting
NTS and VLM neurons were DBH-positive in newborn rats compared with adults,
consistent with reduced DBH immunostaining in the neonatal PVN. Conversely,
fiber immunolabeling for phenylethanolamine N-methyltransferase (PNMT, the
enzyme that converts norepinephrine to epinephrine) within the PVN was most
dense at P1, gradually decreasing to adult-like levels by P21. The dynamic temporal
shift in the density of immunolabeling for these catecholaminergic synthetic
enzymes reflects a gradual maturation of noradrenergic and adrenergic signaling to
the hypothalamus in rats during the first 3 weeks of life (Rinaman 2001). It remains
unclear whether the shift in immunolabeling patterns represents parallel increases
and reductions in axonal terminal density arising from PNMT-negative versus
PNMT-positive NA neurons, or whether it represents altered enzyme content within
individual neurons whose axons innervate the neuroendocrine hypothalamus.
Fig. 13.3 Immunolabeling for yellow fluorescent protein (YFP) expressed under control of the
preproglucagon (PPG) promoter in transgenic PPG-YFP mice. Top panel: YFP immunoperoxidase
labeling of PPG neurons and processes within the caudal NTS in a newborn mouse (P0). Bottom
panel: green YFP immunofluorescence is co-localized with red GLP1 immunofluorescence to
generate yellow double-labeled neurons within the caudal NTS in a 6-day old PPG-YFP mouse
(P6). C: central canal. Scale bar: 50 μm
are expressed by NA neurons within the caudal NTS, which do not express leptin
receptors in mice (Huo et al. 2008). The potential significance of these species
differences in expression of leptin receptors in the caudal NTS remains unclear, but
it presumably is linked to differential effects of leptin on hindbrain-hypothalamic
circuits in both developing and mature animals.
Another study investigated the postnatal maturation of central neural Fos responses
to i.p. injection of LiCl, a malaise-inducing agent (Koehnle and Rinaman 2007).
Compared to Fos activation in neonatal control rats injected with physiological
saline, LiCl did not increase Fos in the PVN or other forebrain regions at P0, but
did so at P7 and later, with maximal PVN Fos responses to LiCl achieved by P14.
Comparable results have been obtained examining central Fos responses to an acute
LPS challenge in developing rats (Oladehin and Blatteis 1995). These findings
provide additional evidence that vagal sensory signaling to the neuroendocrine
hypothalamus undergoes significant functional maturation in rats during the first
2–3 weeks of postnatal life (Rinaman and Koehnle 2010).
Given the apparent inability of systemic CCK, LiCl, or LPS to activate hypotha-
lamic neurons in neonatal rats, one might hypothesize that hypothalamic endocrine
neurons are refractory to all sources of excitatory input early in development. This
360 L. Rinaman
The findings summarized above (Sect. 13.3.5) support the view that ascending
neural pathways conveying visceral sensory signals from hindbrain to hypothalamus
are functionally immature in neonatal rats. Thus, these circuits may exhibit plasticity
as they assemble during a developmentally defined sensitive period. Highly
specialized mechanisms are crucial for the initial establishment of postsynaptic
specializations during synaptogenesis, and for activity-related changes in synaptic
strength that underlie experience-dependent plasticity. By analogy with other CNS
systems, evoked neural activity within ascending visceral sensory pathways may
shape ongoing synapse formation and stabilization of visceral sensory inputs to the
neuroendocrine hypothalamus during the first few weeks of postnatal life.
In rats, the first 2 weeks of postnatal life correspond to the so-called SHRP,
characterized by reduced HPA axis responsiveness to stressful stimuli (Walker
et al. 1991). Most of the peripheral and central components of stress-responsive
circuits are capable of being activated in neonatal rats, given sufficient experimental
or natural conditions (Levine 2001). However, HPA axis responsiveness in neonatal
rats is apparently suppressed by maternal factors that generate interoceptive signals
in their developing offspring, including thermal, chemical, and mechanoreceptive
components (Rinaman and Koehnle 2010).
13 Organization and Postnatal Development of Visceral Sensory Inputs to the. . . 361
Plasma levels of GI-derived CCK and pituitary-derived GH fall when rat pups are
separated from their dam and increase during and after maternal reunion and feeding
(Kuhn et al. 1979; Weller et al. 1992). Experimental paradigms that reduce/fragment
the early life maternal care received by rat pups exert a significant impact on adult
measures of stress-evoked norepinephrine release within the PVN, modulating
stimulus-induced activation of CRH endocrine neurons and anterior pituitary release
of ACTH (Liu et al. 2000; Walker et al. 2017). Further, CCK-1 receptors are
especially abundant in the upper GI tract in newborn rats (Robinson et al. 1987).
Based on the results of experiments in which early life maternal care was
manipulated together with manipulation of CCK-1 receptor signaling, we have
proposed that the impact of maternal care on central visceral circuits is based, at
least in part, on the activity of GI vagal sensory input to the NTS (Weber et al. 2009;
Rinaman and Koehnle 2010), which in turn regulates the neuroendocrine hypothal-
amus through ascending NA, GLP1, and potentially other neurochemical signaling
pathways.
13.5 Perspectives
Key References
Bouret and Simerly (2007)—This review article summarizes evidence that the
intrinsic hypothalamic circuits that integrate interoceptive signals (here, leptin
signaling from visceral adipose tissue) undergo a significant amount of postnatal
362 L. Rinaman
References
Bienkowski MS, Rinaman L (2013) Common and distinct neural inputs to the medial central
nucleus of the amygdala and anterior ventrolateral bed nucleus of stria terminalis in rats.
Brain Struct Funct 218:187–208
Bouret SG, Simerly RB (2007) Development of leptin-sensitive circuits. J Neuroendocrinol
19:575–582
Bouret SG, Draper SJ, Simerly RB (2004) Formation of projection pathways from the arcuate
nucleus of the hypothalamus to hypothalamic regions implicated in the neural control of feeding
behavior in mice. J Neurosci 24:2797–2805
Briski KP, Shrestha PK (2016) Hindbrain estrogen receptor-beta antagonism normalizes reproduc-
tive and counter-regulatory hormone secretion in hypoglycemic steroid-primed ovariectomized
female rats. Neuroscience 331:62–71
Chang HC, Chia KF, Huang JJ, Lim RKS (1939) Vagus-post-pituitary reflex; antidiuretic effect.
Clin J Physiol 14:161–172
Chen C-T, Dun SL, Dun NJ, Chang J-K (1999) Prolactin-releasing peptide-immunoreactivity in A1
and A2 noradrenergic neurons of the rat medulla. Brain Res 822:276–279
13 Organization and Postnatal Development of Visceral Sensory Inputs to the. . . 363
Coyle JT, Axelrod J (1971) Development of the uptake and storage of L-[3H]norepinephrine in the
rat brain. J Neurochem 18:2061–2075
Date Y, Murakami N, Toshinai K, Matsukura S, Niijima A, Matsuo H, Kangawa K, Nakazato M
(2002) The role of the gastric afferent vagal nerve in ghrelin-induced feeding and growth
hormone secretion in rats. Gastroenterology 123:1120–1128
Day TA, Blessing W, Willoughby JO (1980) Noradrenergic and dopaminergic projections to the
medial preoptic area of the rat. A combined horseradish peroxidase/catecholamine fluorescence
study. Brain Res 193:543–548
Day T, Ferguson AV, Renaud L (1985) Noradrenergic afferents facilitate the activity of
tuberoinfundibular neurons of the hypothalamic paraventricular nucleus. Neuroendocrinology
41:17
Farkas I, Vastagh C, Farkas E, Balint F, Skrapits K, Hrabovszky E, Fekete C, Liposits Z (2016)
Glucagon-like peptide-1 excites firing and increases GABAergic miniature postsynaptic
currents (mPSCs) in gonadotropin-releasing hormone (GnRH) neurons of the male mice via
activation of nitric oxide (NO) and suppression of endocannabinoid signaling pathways. Front
Cell Neurosci 10:214
Godement P, Vanselow J, Thanos S, Bonhoeffer F (1987) A study in developing visual systems
with a new method of staining neurones and their processes in fixed tissue. Development
101:697–713
Goke R, Larsen PJ, Mikkelsen JD, Sheikh SP (1995) Distribution of GLP-1 binding sites in the rat
brain: evidence that exendin-4 is a ligand of brain GLP-1 binding sites. Eur J Neurosci
7:2294–2300
Grill HJ, Norgren R (1978) Chronically decerebrate rats demonstrate satiation but not bait shyness.
Science 201:267–269
Heppner KM, Baquero AF, Bennett CM, Lindsley SR, Kirigiti MA, Bennett B, Bosch MA, Mercer
AJ, Ronnekleiv OK, True C, Grove KL, Smith MS (2017) GLP-1R signaling directly activates
arcuate nucleus kisspeptin action in brain slices but does not rescue luteinizing hormone
inhibition in ovariectomized mice during negative energy balance. eNeuro 4(1):
ENEURO.0198-16
Huo L, Gamber KM, Grill HJ, Bjorbaek C (2008) Divergent leptin signaling in proglucagon
neurons of the nucleus of the solitary tract in mice and rats. Endocrinology 149:492–497
Ibrahim BA, Briski KP (2014) Role of dorsal vagal complex A2 noradrenergic neurons in hindbrain
glucoprivic inhibition of the luteinizing hormone surge in the steroid-primed ovariectomized
female rat: effects of 5-thioglucose on A2 functional biomarker and AMPK activity. Neurosci-
ence 269:199–214
Jin Y, Ueta Y, Kannan H, Yamashita H (1993) Synaptic inputs from the stomach to
tuberoinfundibular neurons in the paraventricular nucleus of the hypothalamus in rats. Brain
Res 617:151–154
Kamilaris TC, Johnson EO, Calogero AE, Kalogeras KT, Bernardini R, Chrousos GP, Gold PW
(1992) Cholecystokinin-octapeptide stimulates hypothalamic-pituitary-adrenal function in rats:
role of corticotropin-releasing hormone. Endocrinology 130:1764–1774
Kaminski KL, Watts AG (2012) Intact catecholamine inputs to the forebrain are required for
appropriate regulation of corticotrophin-releasing hormone and vasopressin gene expression
by corticosterone in the rat paraventricular nucleus. J Neuroendocrinol 24:1517–1526
Khachaturian H, Sladek JR (1980) Simultaneous monoaminehistofluorescence and neuropeptide
immunocytochemistry: III. Ontogeny of catecholamine varicosities and neurophysin neurons in
the rat supraoptic and paraventricular nuclei. Peptides 1:77–95
Kinzig KP, D’Alessio DA, Herman JP, Sakai RR, Vahl TP, Figueiredo HF, Murphy EK, Seeley RJ
(2003) CNS glucagon-like peptide-1 receptors mediate endocrine and anxiety responses to
interoceptive and psychogenic stressors. J Neurosci 23:6163–6170
Koehnle T, Rinaman L (2007) Progressive postnatal increases in Fos immunoreactivity in the
forebrain and brainstem of rats after viscerosensory stimulation with lithium chloride. Am J
Physiol Regul Integr Comp Physiol 292:R1212–R1223
364 L. Rinaman
Kreisler AD, Davis EA, Rinaman L (2014) Differential activation of chemically identified neurons
in the caudal nucleus of the solitary tract in non-entrained rats after intake of
satiating vs. non-satiating meals. Physiol Behav 136:47–54
Kuhn CM, Evoniuk G, Schanberg SM (1979) Loss of tissue sensitivity to growth hormone during
maternal deprivation in rats. Life Sci 25:2090–2097
Larsen PJ, Tang-Christensen M, Jessop DS (1997) Central administration of glucagon-like peptide-
1 activates hypothalamic neuroendocrine neurons in the rat. Endocrinology 138:4445–4455
Levine S (2001) Primary social relationships influence the development of the hypothalamic-
pituitary-adrenal axis in the rat. Physiol Behav 73:255–260
Liu D, Caldji C, Sharma S, Plotsky PM, Meaney MJ (2000) Influence of neonatal rearing conditions
on stress-induced adrenocorticotropin responses and norepinephrine release in the hypothalamic
paraventricular nucleus. J Neuroendocrinol 12:5–12
Llewellyn-Smith IJ, Reimann F, Gribble FM, Trapp S (2011) Preproglucagon neurons project
widely to autonomic control areas in the mouse brain. Neuroscience 180:111–121
Maniscalco JW, Kreisler AD, Rinaman L (2013) Satiation and stress-induced hypophagia: examin-
ing the role of hindbrain neurons expressing prolactin-releasing peptide or glucagon-like peptide
1. Front Neurosci 6:199
Menétrey D, dePommery J (1991) Origins of spinal ascending pathways that reach central areas
involved in visceroception and visceronociception in the rat. Eur J Neurosci 3:249–259
Morales T, Hinuma S, Sawchenko PE (2000) Prolactin-releasing peptide is expressed in afferents to
the endocrine hypothalamus, but not in neurosecretory neurones. J Neuroendocrinol
12:131–140
Nelson EE, Alberts JR, Tian Y, Verbalis JG (1998) Oxytocin is elevated in plasma of 10-day-old
rats following gastric distension. Dev Brain Res 111:301–303
Nillni EA (2010) Regulation of the hypothalamic thyrotropin releasing hormone (TRH) neuron by
neuronal and peripheral inputs. Front Neuroendocrinol 31:134–156
Ohkura S, Tanaka T, Nagatani S, Bucholtz DC, Tsukamura H, Maeda K, Foster DL (2000) Central,
but not peripheral, glucose-sensing mechanisms mediate glucoprivic suppression of pulsatile
luteinizing hormone secretion in the sheep. Endocrinology 141:4472–4480
Oladehin A, Blatteis CM (1995) Lipopolysaccharide-induced fos expression in hypothalamic nuclei
of neonatal rats. Neuroimmunomodulation 2:282–289
Palkovits M (1999) Interconnections between the neuroendocrine hypothalamus and the central
autonomic system. Front Neuroendocrinol 20:270–295
Plotsky P (1987) Facilitation of immunoreactive corticotropin-releasing factor secretion into the
hypophysial-portal circulation after activation of catecholaminergic pathways or central norepi-
nephrine injection. Endocrinology 121:924–930
Raybould HE, Tache Y (1988) Cholecystokinin inhibits gastric motility and emptying via a
capsaicin-sensitive vagal pathway in rats. Am J Phys 255:G242–G246
Rinaman L (1999a) Interoceptive stress activates glucagon-like peptide-1 neurons that project to the
hypothalamus. Am J Phys 277:R582–R590
Rinaman L (1999b) A functional role for central glucagon-like peptide-1 receptors in lithium
chloride-induced anorexia. Am J Phys 277:R1537–R1540
Rinaman L (2001) Postnatal development of catecholamine inputs to the paraventricular nucleus of
the hypothalamus in rats. J Comp Neurol 438:411–422
Rinaman L (2007) Visceral sensory inputs to the endocrine hypothalamus. Front Neuroendocrinol
28:50–60
Rinaman L (2010) Hindbrain noradrenergic A2 neurons: diverse roles in autonomic, endocrine,
cognitive, and behavioral functions. Am J Physiol Regul Integr Comp Physiol 00556:02010
Rinaman L, Koehnle TJ (2010) Development of central visceral circuits. In: Blumberg MS,
Freeman JH, Robinson SR (eds) Oxford handbook of developmental behavioral neuroscience.
Oxford University Press, New York, pp 298–321
Rinaman L, Levitt P (1993) Establishment of vagal sensorimotor circuits during fetal development
in rats. J Neurobiol 24:641–659
13 Organization and Postnatal Development of Visceral Sensory Inputs to the. . . 365
Rinaman L, Hoffman GE, Stricker EM, Verbalis JG (1994) Exogenous cholecystokinin activates
cFos expression in medullary but not hypothalamic neurons in neonatal rats. Dev Brain Res
77:140–145
Rinaman L, Hoffman GE, Dohanics J, Le WW, Stricker EM, Verbalis JG (1995) Cholecystokinin
activates catecholaminergic neurons in the caudal medulla that innervate the paraventricular
nucleus of the hypothalamus in rats. J Comp Neurol 360:246–256
Rinaman L, Stricker EM, Hoffman GE, Verbalis JG (1997) Central c-fos expression in neonatal and
adult rats after subcutaneous injection of hypertonic saline. Neuroscience 79:1165–1175
Robinson PH, Moran TH, Goldrich M, McHugh PR (1987) Development of cholecystokinin
binding sites in the rat upper gastrointestinal tract. Am J Phys 252:G529–G534
Robinson PH, Moran TH, McHugh PR (1988) Cholecystokinin inhibits independent ingestion in
neonatal rats. Am J Phys 255:R14–R20
Roman CW, Sloat SR, Palmiter RD (2017) A tale of two circuits: CCK NTS neuron stimulation
controls appetite and induces opposing motivational states by projections to distinct brain
regions. Neuroscience 358:316–324
Saper CB, Loewy AD (1980) Efferent connections of the parabrachial nucleus in the rat. Brain Res
197:291–317
Sawchenko PE, Swanson LW (1982) The organization of noradrenergic pathways from the
brainstem to the paraventricular and supraoptic nuclei in the rat. Brain Res Rev 4:275–325
Shapiro RE, Miselis RR (1985) The central neural connections of the area postrema of the rat.
J Comp Neurol 234:344–364
Shughrue PJ, Lane MV, Merchenthaler I (1996) Glucagon-like peptide-1 receptor (GLP1-R)
mRNA in the rat hypothalamus. Endocrinology 137:5159–5162
Stornetta RL, Sevigny CP, Guyenet PG (2002) Vesicular glutamate transporter DNPI/VGLUT2
mRNA is present in C1 and several other groups of brainstem catecholaminergic neurons.
J Comp Neurol 444:191–206
Szawka RE, Poletini MO, Leite CM, Bernuci MP, Kalil B, Mendonca LB, Carolino RO, Helena
CV, Bertram R, Franci CR, Anselmo-Franci JA (2013) Release of norepinephrine in the
preoptic area activates anteroventral periventricular nucleus neurons and stimulates the surge
of luteinizing hormone. Endocrinology 154:363–374
Thieblot L, Duchene-Marullaz P, Berthelay J (1957) Nouvelle preuve de la libération de principes
posthypophysaires par l’excitation centripète du vague. Ann Endocrinol 18:651–653
Ueta Y, Kannan H, Higuchi T, Negoro H, Yamashita H (1993) CCK-8 excites oxytocin-secreting
neurons in the paraventricular nucleus in rats - possible involvement of noradrenergic pathway.
Brain Res Bull 32:453–459
Verbalis JG, McCann MJ, McHale CM, Stricker EM (1986) Oxytocin secretion in response to
cholecystokinin and food: differentiation of nausea from satiety. Science 232:1417–1419
Verbalis JG, Richardson DW, Stricker EM (1987) Vasopressin release in response to nausea-
producing agents and cholecystokinin in monkeys. Am J Phys 252:R749–R753
Verbalis JG, Stricker EM, Robinson AG, Hoffman GE (1991) Cholecystokinin activates cFos
expression in hypothalamic oxytocin and corticotropin-releasing hormone neurons.
J Neuroendocrinol 3:205–213
Walker CD, Scribner VA, Cascio CS, Dallman MF (1991) The pituitary-adrenocortical system of
neonatal rats is responsive to stress throughout development in a time dependent and stress-
specific fashion. Endocrinology 128:1385–1396
Walker CD, Bath KG, Joels M, Korosi A, Larauche M, Lucassen PJ, Morris MJ, Raineki C, Roth
TL, Sullivan RM, Tache Y, Baram TZ (2017) Chronic early life stress induced by limited
bedding and nesting (LBN) material in rodents: critical considerations of methodology,
outcomes and translational potential. Stress 20:421–448
Weber BC, Manfredo HN, Rinaman L (2009) A potential gastrointestinal link between enhanced
postnatal maternal care and reduced anxiety-like behavior in adolescent rats. Behav Neurosci
123:1178–1184
366 L. Rinaman
Weller A, Corp ES, Tyrka A, Ritter RC, Brenner L, Gibbs J, Smith GP (1992) Trypsin inhibitor and
maternal Reunion increase plasma cholecystokinin in neonatal rats. Peptides 13:939–941
Yano T, Iijima N, Kataoka Y, Hinuma S, Tanaka M, Ibata Y (2001) Developmental expression of
prolactin releasing peptide in the rat brain: localization of messenger ribonucleic acid and
immunoreactive neurons. Dev Brain Res 128:101–111
Zheng H, Stornetta RL, Agassandian K, Rinaman L (2015) Glutamatergic phenotype of glucagon-
like peptide 1 neurons in the caudal nucleus of the solitary tract in rats. Brain Struct Funct
220:3011–3022
Zueco JA, Esquifino AI, Chowen JA, Alvarez E, Castrillon PO, Blazquez E (1999) Coexpression of
glucagon-like peptide-1 (GLP-1) receptor, vasopressin, and oxytocin mRNAs in neurons of the
rat hypothalamic supraoptic and paraventricular nuclei: effect of GLP-1(7-36)amide on vaso-
pressin and oxytocin release. J Neurochem 72:10–16
Astrocytes and Development
of Neuroendocrine Circuits 14
Lydia L. DonCarlos and Julie A. Chowen
Abstract
Astrocytes, one of the most abundant cell types in the hypothalamus, perform a
myriad of functions throughout all stages of development. Radial glia, astroglia-
like progenitor cells, are fundamental for neurogenesis, neuronal migration, axon
extension, and synaptic formation throughout the brain. These progenitor cells
also give rise to mature astrocytes, which are essential for the proper functioning
of the brain at all ages. In the hypothalamus, astrocytes play an important role in
all neuroendocrine systems. Together with tanycytes, specialized glial cells
surrounding the third ventricle of the hypothalamus, astrocytes maintain some
of their developmental potential into adulthood, including their progenitor com-
petence and ability to modulate synaptic plasticity. In this chapter, we will
address what is known regarding the role of astrocytes in hypothalamic develop-
ment. We will briefly cover some of the functions that astrocytes perform in
specific neuroendocrine systems, as well as how the maintenance of certain
developmental capacities during later stages of life is fundamental for systemic
adaptation and homeostatic control. Although the roles of astrocytes in neuroen-
docrine control have received increasing attention in recent years, there is still
much to be learned regarding these fascinating cells. Identification of functional
astrocytic populations with specific markers will allow these cells to be targeted
L. L. DonCarlos
Department of Cell and Molecular Physiology, Stritch School of Medicine, Loyola University
Chicago, Maywood, IL, USA
J. A. Chowen (*)
Department of Pediatrics and Pediatric Endocrinology, Hospital Infantil Universitario Niño Jesús,
Princesa Research Institute La Princesa, Madrid, Spain
Centro de Investigación Biomédica en Red de Fisiopatología de la Obesidad y Nutrición
(CIBEROBN), Instituto de Salud Carlos III, IMDEA Food Institute, CEIUAM+CSIC, Madrid,
Spain
e-mail: [email protected]
and more easily manipulated, so that we can begin to learn details about how
astroglia participate in the development of specific hypothalamic circuits.
Keywords
Arcuate nucleus · Astrocytes · Glia · Gliogenesis · Neuroendocrine circuits ·
Paraventricular nucleus · Tanycytes · Suprachiasmatic nucleus · Supraoptic
nucleus
14.1 Introduction
Astrocytes are among the most abundant glial cells in the central nervous system,
with the proportion of astrocytes to other glial cells or to neurons varying depending
on the brain region (Ero et al. 2018). Although astrocytes were early on recognized
as having processes contacting multiple different neurons and communicating with
each other as well (Fig. 14.1), they were relegated to a supporting role for some
years. Research beginning in the 1970s led to increasing recognition of the promi-
nent role of astrocytes in the function of the mature and damaged nervous system,
and, for the purpose of this review, as being key to the development and honing of
neural circuits (Allen and Lyons 2018) including the hypothalamus and neuroendo-
crine systems.
Because they do not generate action potentials, many aspects of astrocytic action
have been difficult to study, and research into their function has often been restricted
by their physiologically inaccessible nature. Overall, astrocytes typically develop
and mature later than neurons, rendering it difficult to study them independent of
developing neuronal circuits. Therefore, our understanding and characterization of
the development and diversity of astrocytes, and how that diversity is determined, is
still in early stages relative to what we know about neurons. However, among the
most informative studies on astrocyte function and plasticity are those that have
examined astrocytes in the hypothalamus, both in development and in response to
manipulations of, for example, gonadal steroid hormones or metabolism. Moreover,
although underappreciated by neuroscientists in general, these studies of astrocytes
in the developing hypothalamus have been both particularly illuminating and yet
particularly enigmatic, as discussed in more detail below.
Glial cells include astrocytes, microglia, oligodendrocytes, tanycytes, and NG2 cells.
They are distinguished by expression of specific proteins and based on their mor-
phology. This group of cells is involved in many of the aspects of brain development
that follow early patterning of the nervous system, including neurogenesis, neuronal
migration, and synaptogenesis (Allen and Lyons 2018; Bosworth and Allen 2017).
These cells are intimately involved in neuroprotection (Belanger and Magistretti
14 Astrocytes and Development of Neuroendocrine Circuits 369
Fig. 14.1 Drawing of astrocytes by Santiago Ramón y Cajal. Courtesy of the Cajal Institute, Cajal
Legacy, Spanish National Research Council (CSIC), Madrid, reprinted with permission. This
detailed drawing by Santiago Ramón y Cajal demonstrates how his keen eye observed the intricate
relationship between astrocytes and neurons. Astrocytes (A) can ensheath neuronal cell bodies (B)
with their projections, as well as contacting neuronal processes and cell bodies (a, b) as well as
blood vessels (c)
2009), neurotransmission (Chung et al. 2015; Perea et al. 2009), and adaptation of
the CNS to physiological changes by participating in synaptic reorganization and
modifications in synaptic function (Chowen et al. 2016; Hatton 1997; Sims et al.
2015). They also sustain neurons and neuronal circuits by supplying nutrients and
maintaining extracellular homeostasis (Leloup et al. 2015; Magistretti and Allaman
2018) with intimate contact and communication between astrocytes and neurons
throughout development.
370 L. L. DonCarlos and J. A. Chowen
Neurons
Neuroblasts
Pluripotent Cell
Astrocytes
Radial glia
Fig. 14.2 Schematic representation of cell lines for development of astrocytes and neurons.
Astrocytes are derived from radial glia
the arcuate and supraoptic nuclei (Kim et al. 2016), indicating that both their number
and proliferation increase.
The first GFAP-positive cells in the suprachiasmatic nucleus (SCN) are found on
E15, suggesting the appearance of maturing astrocytes. The number of GFAP-
positive glia increases until P21, when an adult-like pattern of GFAP immunoreac-
tivity is observed (Botchkina and Morin 1995). In the ventrolateral portion of the
SCN, Munekawa and colleagues (Munekawa et al. 2000) reported that GFAP-
positive cells appear at embryonic day 20, increase at PND0, and then rapidly
increase between PND3 and PND4. These developmental changes are influenced
by light, with neuronal inputs from the retinohypothalamic tract being fundamental.
Although GFAP-positive cells are present in the ependymal layer of the arcuate
nucleus by the first postnatal week, consistent with these cells being developing
tanycytes, it is not until the second postnatal week when some GFAP-positive cells
are also located in more lateral regions of the arcuate nucleus, where astrocytes are
known to reside. Later, around the sixth postnatal week, GFAP-positive cells are
found close to arcuate neurons and the number of these cells increases through the
eighth postnatal week (Suarez et al. 1987). This study suggests that in the arcuate
nucleus, GFAP-positive cells that comprise the astrocytic population do not have a
mature cellular organization until after the eighth week of postnatal life. Although
gliogenesis in the hypothalamus (i.e., the proliferation of astrocytes, in this case)
continues through puberty and adulthood (Ahmed et al. 2008; Mohr et al. 2016), the
rate of incorporation decreases with age.
Postmitotic maturation of astrocytes includes the onset of, and an increase in,
expression of proteins important for astrocytic functions (e.g., glutamate
transporters) and increased arborization of glial projections. Maturation of astrocytes
is suggested to occur, at least in part, in response to surrounding neuronal activity,
which could be involved in the differential phenotype/functionality of astrocytes in
distinct brain regions. Maturation of astrocytes in areas such as the hippocampus
(Bushong et al. 2004) and cortex (Morel et al. 2014) has been shown to occur from
early postnatal life until approximately one month of age. Postnatal maturation of
astrocytes has been documented in some hypothalamic nuclei, but much is yet to be
learned regarding this process in the hypothalamus as it appears to differ to some
extent from that in other brain areas. For example, in the mouse cerebral cortex,
arborization of astrocytes increases between PND4 and PND26, with this rise being
most significant between PND14 and 26 (Morel et al. 2014). In contrast, in the
ventromedial or dorsomedial nuclei of the hypothalamus, this modification in
astrocyte morphology was not observed during this time period. Moreover, loss of
synaptic inputs from vesicular glutamate transporter 1 (VGluT1) positive neurons
during development had no effect on astrocyte arborization in these hypothalamic
nuclei, while in the cortex astrocyte morphology was altered and resulted in a
reduction in the glial ensheathment of synapses in this brain area, suggesting that
372 L. L. DonCarlos and J. A. Chowen
New transcriptional profiling methods are paving the way for a more precise
study of the developmental changes and physiologic responses in potentially hetero-
geneous populations of astrocytes in the neuroendocrine system. One such method is
Act-seq, a modified single cell RNA-seq method of high throughput transcriptional
profiling (Wu et al. 2017). Act-seq was designed to prevent the artificial upregulation
of immediate early genes (IEGs) that typically occurs in response to proteases used
during conventional cell dissociation. The method employs the transcription inhibi-
tor, actinomycin-D, along with modifications in dissociation methods, to inhibit
transcription during cell dissociation prior to running a single cell RNA-seq assay
and cluster analysis. Three distinct clusters of astrocytes were found in the medial
amygdala. While all three clusters expressed known astrocytic markers, each cluster
also had a distinct transcriptional profile, suggestive of unique functions. By
suppressing dissociation-related upregulation of IEGs, the group was able to show
that the three subsets of astrocytes in the medial amygdala differentially upregulate
IEGs during a physiologic response, in this case, drug-induced seizure.
Similar approaches would be advantageous in studying hypothalamic astrocytes,
in particular because the data obtained would assist the identification of specific
genes of interest in neuroendocrine studies, allowing a better understanding of
astrocyte heterogeneity and what circumstances drive that heterogeneity. These
studies would further provide information to design and use additional methods to
study astrocytes, such as translating ribosome affinity purification (TRAP) for
translational profiling of the different subsets of astrocytes, or regional and morpho-
logical studies following environmental, genetic, or pharmacological manipulations.
In a study by Morel et al. (2017) TRAP analysis followed by RNA-Seq on samples
of somatosensory cortex, hippocampus, thalamus, hypothalamus, nucleus
accumbens, and the caudate-putamen identified both a dorsoventral gradient and
regional differences in expression of translating mRNAs. Hypothalamic astrocytes
were reported to most closely resemble thalamic astrocytes, with a 60% overall
similarity in transcriptional profile, and the transcriptional profiles of astrocytes in
these diencephalic regions were different from those in the telencephalic regions
sampled. It will be of great interest to determine the similarities, or differences,
between astrocytes from specific hypothalamic nuclei under physiologically relevant
paradigms. In the next section, we review the actions of astrocytes in the regulation
of neuroendocrine function, and then discuss what is known about the participation
of astrocytes in the development of neuroendocrine systems, and what molecular
mechanisms mediate these effects.
(continued)
14 Astrocytes and Development of Neuroendocrine Circuits 375
It has been four decades since hypothalamic astrocytes were first reported to
participate in the response to hormones and the control of specific neuroendocrine
functions (Tweedle and Hatton 1977). These early studies indicated that glial cells
were involved in the neuroendocrine control of osmotic homeostasis, lactation, and
reproduction (Olmos et al. 1989; Perlmutter et al. 1985; Theodosis et al. 1986;
Tweedle and Hatton 1977). Despite these intriguing early observations, glial cells
have remained in the background in neuroendocrine investigation for many years. It
is only recently that they have begun to receive their due attention, and important
advances have been made in understanding how they affect the functioning of
neuroendocrine systems, although there is still much to be learned. Ultimately,
astrocyte development in the preoptic area and hypothalamus may provide a power-
ful model for future investigation into the development of heterogeneity in astrocytes
in general.
376 L. L. DonCarlos and J. A. Chowen
Soma
Soma
Unsmulated
Soma
Soma
Smulated
Fig. 14.3 Schematic representation of magnocellular neurons of the paraventricular and supraoptic
nucleus. Under basal or unstimulated conditions, the somas (blue) of these neurons are separated by
astrocytic processes (green). When stimulated (e.g., by oxytocin for oxytocin neurons), astrocytic
processes are retracted and contact between the plasma membranes of neurons is increased
Modifications in the physical interactions between neurons and glia are also
involved in the neuroendocrine control of reproduction. Astrocytes in the preoptic
area and arcuate nucleus of the hypothalamus coordinate physiological responses to
estradiol (Garcia-Segura and McCarthy 2004), with these astrocytic responses
regulating the surge of gonadotropin hormone-releasing hormone (GnRH) that
precipitates female ovulation, a function critical for reproduction.
Ovariectomy of nonhuman primates increases the glial coverage of GnRH
neurons in both the preoptic area and medial-basal hypothalamus. This increased
neuronal ensheathment is associated with a decline in the number of synaptic
contacts onto these neurons (Witkin et al. 1991). Gonadal steroid replacement was
shown to at least partially reverse these changes in hypothalamic cellular structure
(Witkin et al. 1991), which agrees with the observation that neuronal-astrocyte
structural organization in the hypothalamus is modified throughout the estrous
cycle in relationship with changes in gonadal steroid levels.
In rodents, hypothalamic astrocytes also undergo morphological modifications in
association with the physiological hormonal variations that occur throughout the
estrous cycle and this is associated with coordinated changes in the number of
synapses (Fernandez-Galaz et al. 1997; Garcia-Segura et al. 2008). Between the
morning and afternoon of proestrus, astroglial processes increase their coverage of
neuronal cell surfaces and GABAergic synaptic terminals are removed from the
soma of arcuate neurons, and this coincides with the luteinizing hormone surge
(Csakvari et al. 2008). These glial processes remain in place, separating the presyn-
aptic terminals from the postsynaptic membrane until the morning of estrus. At this
time the glial processes retract so that by the morning of metestrus, GABAergic
synaptic contacts are reestablished in the arcuate nucleus (Csakvari et al. 2007).
14 Astrocytes and Development of Neuroendocrine Circuits 379
During proestrus when GABAergic inputs to the cell soma are decreased there is a
rise in the number of excitatory synapses on dendritic spines and together this results
in increased firing of arcuate neurons (Csakvari et al. 2007). The synaptic changes
observed in the arcuate nucleus during the estrous cycle are most likely driven by the
proestrus rise in circulating 17β-estradiol levels, as administration of this sex steroid
to ovariectomized rats can mimic the changes observed in cycling rats (Naftolin et al.
1996). A subpopulation of neurons in the arcuate nucleus projects to the median
eminence and controls pituitary hormone secretion and this estrogen-induced synap-
tic remodeling most likely induces functional changes in some neurosecretory
neurons to modulate pituitary hormone secretion (Csakvari et al. 2008), as well as
reorganization of synaptic circuits that control GnRH neurons (Garcia-Segura et al.
2008) and those involved in metabolic control (Gao et al. 2007).
The astrocytic responses in the preoptic area and arcuate nucleus are fascinating
because the estradiol-induced plasticity in morphology of astrocytic processes is
entirely different in the preoptic area from that in the arcuate nucleus, and yet
completely coordinated in ensuring appropriate regulation of female reproduction
(Garcia-Segura and McCarthy 2004). Astrocytes in both regions respond to estradiol
with alterations in extension of processes that affect neurotransmission, but in the
preoptic area, astrocytic processes respond by retracting processes that are
interposed between excitatory glutamatergic inputs and GnRH neurons, stimulating
GnRH release. In contrast, the astrocytic processes in the arcuate nucleus respond to
estradiol with interposition of processes that cut off inhibitory GABAergic synapses,
and further encourage GnRH release in response to high levels of estradiol and then
progesterone. The result is a coordinated increase in excitatory inputs and decrease
in inhibitory inputs to GnRH neurons that supports a sustained and powerful release
of GnRH. But, many questions remain. For example, how do the astrocytes in two
different regions develop the perfectly coordinated, but opposite, morphological
responses to the same stimulus? Does this differential response depend on the
developmental environment or the neuronal environment, since the preoptic area is
derived from telencephalic anlage, whereas the arcuate nucleus is derived from the
diencephalic anlage? How does local heterogeneity of astrocyte gene expression
develop, and how does this affect function?
Evidence indicates that the developmental environment can alter the number of
axosomatic synapses in the arcuate nucleus. In rodents, the number of axosomatic
synapses increases between postnatal days 10 and 20 in both males and females
(Perez et al. 1990). However, from postnatal day 20 onward females have more
axosomatic synapses than males, with this sex difference at least partially dependent
on the neonatal sex steroid environment (Perez et al. 1990). Astrocyte morphology
and expression of GFAP are also modified by the sex steroid environment during this
responsive period of hypothalamic development (Chowen et al. 1995), allowing
astrocytes to intervene in critical synaptic interactions (Garcia-Segura et al. 1995).
380 L. L. DonCarlos and J. A. Chowen
14.4.3 Metabolism
Maternal obesity
Neonatal overnutrion
Lepn
BDNF
IL6
Increased
number of
astrocytes
3rd ventricle
Arcuate nucleus
Fig. 14.4 Factors changing the number of astrocytes in the arcuate nucleus during fetal and
neonatal development. Several different physiological signals can alter astrocyte numbers at
different times in development. BDNF Brain derived neurotrophic factor, IL6 Interleukin 6
Many questions remain regarding the interaction of astrocytes and neurons and
how they control appetite and systemic metabolism. For example, astrocytic contact
with both POMC and NPY neurons is increased in obese subjects, in association
with a decrease in the number of somatic synaptic inputs in both populations of
neurons (Horvath et al. 2010). However, on POMC neurons this decrease is due to a
decline in inhibitory inputs, whereas on NPY neurons it is the result of a reduction in
stimulatory inputs, which together promote an overall balance toward an anorexi-
genic signal. How is this specific rewiring coordinated? What is the mechanism of
cross talk between astroglia and neurons that generates this specific synaptic
rewiring, both during development and in adults? Are the astrocytes that surround
these two neuronal populations different, even though they both respond to leptin? Is
there cross talk between astrocytes and neurons during development that ensures that
the required functional circuits in each hypothalamic area are formed, although the
signals involved remain elusive? The hypothalamic circuits controlling each neuro-
endocrine system or axis are often studied and spoken of as if they function
independently, but there must also be precise communication between these circuits
to maintain systemic homeostasis. Astrocytes and radial glia are involved in the
generation of neurons and their projections. How they determine the precise organi-
zation of the neuroendocrine hypothalamus remains largely unknown.
Studies using genetically modified mice can be used to provide evidence for how
neurons affect astrocytes. For example, Gnasxl-deficient mice, which lack a
G-protein alpha stimulatory subunit that regulates energy balance, are lean and
hypermetabolic (Holmes et al. 2016). In these animals, expression of GFAP in the
adult hypothalamus was greatly diminished and the numbers of astrocytes and a
subpopulation of tanycytes (alpha) were also diminished. In juvenile animals as late
as postnatal day 15, however, the numbers of astrocytes and tanycytes were not
different in mutants. Since Gnasxl is expressed only in neurons, this study provides
tentative evidence that late development of astrocytes in the hypothalamus is
modulated by neuronal activity, although it is also possible that any effect of these
mutants on astrocytes may be secondary to undernutrition.
Tanycytes, specialized radial glia located in the ventral wall of the third ventricle,
have been proposed to be a source of neural progenitor cells in the adult hypothala-
mus (Haan et al. 2013; Lee et al. 2012; Robins et al. 2013). Tanycytes are classified
into subtypes depending on their location and function, with some studies indicating
that all subtypes have neural progenitor properties. However, other studies suggest
that α2-tanycytes generate β-tanycytes in vivo and that only α2-tanycytes are
capable of generating proliferative neurospheres in vitro (Robins et al. 2013).
Tanycytes have also been proposed to participate in the neuroendocrine control of
reproduction and metabolism (Ojeda et al. 2003; Prevot et al. 2018) and to modulate
hypothalamic control of the thyroid axis (Muller-Fielitz et al. 2017). Tanycytes have
end feet that extend into the VMH, arcuate nucleus, and median eminence
384 L. L. DonCarlos and J. A. Chowen
(Rodriguez et al. 2005) and can be retracted. They contribute to the regulation of
GnRH release, as their retractable end feet are interlaced with GnRH synaptic
terminals and the portal vasculature in the median eminence (Baroncini et al.
2007; Ojeda et al. 2008; Prevot et al. 2018). Extension of tanycytic processes to
ensheath GnRH terminals prevents release of GnRH into the portal vasculature.
These tanycytic processes retract during the preovulatory stage of the estrous cycle,
which permits the contact of GnRH terminals with portal capillaries for neuropeptide
release (Bellefontaine et al. 2011; Prevot et al. 1999). More than 30 years ago, Sergio
Ojeda hypothesized that glial cells released substances that modulated the release of
GnRH to the portal vasculature (Ojeda 1994). These signals are now known to
include prostaglandin E2, transforming growth factor beta, and nitric oxide (NO),
which stimulate and coordinate GnRH neuronal firing (Bellefontaine et al. 2011;
Clasadonte et al. 2011; Melcangi et al. 1995). More recently, tanycytic were shown
to produce semaphorin7A which binds to plexinC1 and β1-integrin (Parkash et al.
2015). The levels of semaphorin7A are modulated by ovarian steroids, and the
signaling of semaphorin 7A is necessary for normal estrous cycling and fertility. It
is proposed that semaphorin7A acts through is receptors in a coordinated fashion to
induce expansion of tanycytic end feet and promote the retraction of GnRH nerve
terminals from the pericapillary space (Parkash et al. 2015).
14.7 Perspectives
In sum, astrocytes have been difficult to study because: (1) they are not electrically
active and do not generate action potentials; (2) they have divergent morphologies
(including small nuclei and irregular processes); and (3) there is a lack of specific
antibodies or other markers that identify astrocytic subpopulations.
The lack of astrocyte-specific reagents is certainly problematic. For example,
GFAP is found in astrocytes and progenitor cells and not all astrocytes express
GFAP. Moreover, GFAP immunoreactivity does not allow full visualization of
astrocyte morphology, and the extent of GFAP staining within astrocytes may be
less indicative of morphological changes than of changes in GFAP expression. Other
difficulties relate to the problem of visualizing astrocytes, especially at the same time
as neurons since they are often found in different planes, making it difficult to
simultaneously label and visualize the nuclei of both cell types.
Controversies in studies of astrocytes have also arisen recently, even for issues
that were previously considered to be somewhat resolved. For example, the idea that
there are co-transmitters released by astrocytes that assist neuronal neurotransmis-
sion has been called into question (Wolosker et al. 2016, 2017). Does the concept of
the tripartite synapse hold true, or is this functionally restricted to developmental
stages? Are calcium waves between astrocytes an artifact of culture conditions,
rather than an in vivo event? Should the umbrella term “astrocyte” be eliminated
in favor of more specific terms that refer to distinct subtypes of this class of glia? Do
astrocytes deliver glucose, lactate or both to neurons?
14 Astrocytes and Development of Neuroendocrine Circuits 385
Key References
Ahmed et al. (2008). Challanged the dogma of that pubertal hormones, as opposed to
sex steroids during embryological/neonatal development did not affect
neurogenesis.
Bouret et al. (2004). Seminal study in understanding how early metabolic/hormonal
changes can have long-term effects on metabolism.
Malatestaet al. (2003). Important contributions in order to understand radial glia.
Mong et al. (1996). One of the first studies to demonstrate sex differences in glial
cells in the hypothalamus.
Morel et al. (2017). Important advances in understanding differences in astrocytes in
different brain regions.
Naftolin et al. (1996). An early study associating synaptic remodeling in the hypo-
thalamus with a physiological outcome.
Perlmutter et al. (1985). A seminal study in understanding the physiological out-
come of interactions between neurons and glia in the hypothalamus.
Pfrieger et al. (1997). A clear advance in undestanding the role of astrocytes in
synaptic transmission.
Prevot et al. (2018). A recent update on the implications of tanycytes in neuroendo-
crine control.
Theodosis et al. (1986). A seminal study of neuron-glial interactions in physiology.
386 L. L. DonCarlos and J. A. Chowen
References
Ahima RS, Bjorbaek C, Osei S, Flier JS (1999) Regulation of neuronal and glial proteins by leptin:
implications for brain development. Endocrinology 140:2755–2762
Ahmed EI, Zehr JL, Schulz KM, Lorenz BH, DonCarlos LL, Sisk CL (2008) Pubertal hormones
modulate the addition of new cells to sexually dimorphic brain regions. Nat Neurosci
11:995–997
Allen NJ, Lyons DA (2018) Glia as architects of central nervous system formation and function.
Science 362:181–185
Altman J, Bayer SA (1986) The development of the rat hypothalamus. Adv Anat Embryol Cell Biol
100:1–178
Amateau SK, McCarthy MM (2002) Sexual differentiation of astrocyte morphology in the devel-
oping rat preoptic area. J Neuroendocrinol 14:904–910
Argente-Arizon P, Ros P, Diaz F, Fuente-Martin E, Castro-Gonzalez D, Sanchez-Garrido MA,
Barrios V, Tena-Sempere M, Argente J, Chowen JA (2016) Age and sex dependent effects of
early overnutrition on metabolic parameters and the role of neonatal androgens. Biol Sex Differ
7:26
Baroncini M, Allet C, Leroy D, Beauvillain JC, Francke JP, Prevot V (2007) Morphological
evidence for direct interaction between gonadotrophin-releasing hormone neurones and
astroglial cells in the human hypothalamus. J Neuroendocrinol 19:691–702
Bayer SA, Altman J (1987) Development of the preoptic area: time and site of origin, migratory
routes, and settling patterns of its neurons. J Comp Neurol 265:65–95
Belanger M, Magistretti PJ (2009) The role of astroglia in neuroprotection. Dialog Clin Neurosci
11:281–295
Bellefontaine N, Hanchate NK, Parkash J, Campagne C, de Seranno S, Clasadonte J, d’Anglemont
de Tassigny X, Prevot V (2011) Nitric oxide as key mediator of neuron-to-neuron and
endothelia-to-glia communication involved in the neuroendocrine control of reproduction.
Neuroendocrinology 93:74–89
Bonfanti L, Poulain DA, Theodosis DT (1993) Radial glia-like cells in the supraoptic nucleus of the
adult rat. J Neuroendocrinol 5:1–5
Bonfardin VD, Fossat P, Theodosis DT, Oliet SH (2010) Glia-dependent switch of kainate receptor
presynaptic action. J Neurosci 30:985–995
Borg ML, Reichenbach A, Lemus M, Oldfield BJ, Andrews ZB, Watt MJ (2016) Central adminis-
tration of the ciliary neurotrophic factor analogue, axokine, does not play a role in long-term
energy homeostasis in adult mice. Neuroendocrinology 103:223–229
Boston BA, Blaydon KM, Varnerin J, Cone RD (1997) Independent and additive effects of central
POMC and leptin pathways on murine obesity. Science 278:1641–1644
Bosworth AP, Allen NJ (2017) The diverse actions of astrocytes during synaptic development. Curr
Opin Neurobiol 47:38–43
Botchkina GI, Morin LP (1995) Ontogeny of radial glia, astrocytes and vasoactive intestinal peptide
immunoreactive neurons in hamster suprachiasmatic nucleus. Brain Res Dev Brain Res
86:48–56
Bouret SG, Simerly RB (2007) Development of leptin-sensitive circuits. J Neuroendocrinol 19
(8):575–582
Bouret SG, Draper SJ, Simerly RB (2004) Trophic action of leptin on hypothalamic neurons that
regulate feeding. Science 304:108–110
Bourque CW, Oliet SH (1997) Osmoreceptors in the central nervous system. Annu Rev Physiol
59:601–619
Bourque CW, Oliet SH, Richard D (1994) Osmoreceptors, osmoreception, and osmoregulation.
Front Neuroendocrinol 15:231–274
Bushong EA, Martone ME, Ellisman MH (2004) Maturation of astrocyte morphology and the
establishment of astrocyte domains during postnatal hippocampal development. Int J Dev
Neurosci 22:73–86
14 Astrocytes and Development of Neuroendocrine Circuits 387
Caruso C, Carniglia L, Durand D, Gonzalez PV, Scimonelli TN, Lasaga M (2012) Melanocortin
4 receptor activation induces brain-derived neurotrophic factor expression in rat astrocytes
through cyclic AMP-protein kinase A pathway. Mol Cell Endocrinol 348:47–54
Chowen JA, Busiguina S, Garcia-Segura LM (1995) Sexual dimorphism and sex steroid modula-
tion of glial fibrillary acidic protein messenger RNA and immunoreactivity levels in the rat
hypothalamus. Neuroscience 69:519–532
Chowen JA, Argente-Arizon P, Freire-Regatillo A, Frago LM, Horvath TL, Argente J (2016) The
role of astrocytes in the hypothalamic response and adaptation to metabolic signals. Prog
Neurobiol 144:68–87
Chung WS, Welsh CA, Barres BA, Stevens B (2015) Do glia drive synaptic and cognitive
impairment in disease? Nat Neurosci 18:1539–1545
Clasadonte J, Poulain P, Hanchate NK, Corfas G, Ojeda SR, Prevot V (2011) Prostaglandin E2
release from astrocytes triggers gonadotropin-releasing hormone (GnRH) neuron firing via EP2
receptor activation. Proc Natl Acad Sci U S A 108:16104–16109
Collado P, Beyer C, Hutchison JB, Holman SD (1995) Hypothalamic distribution of astrocytes is
gender-related in Mongolian gerbils. Neurosci Lett 184:86–89
Csakvari E, Hoyk Z, Gyenes A, Garcia-Ovejero D, Garcia-Segura LM, Parducz A (2007) Fluctua-
tion of synapse density in the arcuate nucleus during the estrous cycle. Neuroscience
144:1288–1292
Csakvari E, Kurunczi A, Hoyk Z, Gyenes A, Naftolin F, Parducz A (2008) Estradiol-induced
synaptic remodeling of tyrosine hydroxylase immunopositive neurons in the rat arcuate nucleus.
Endocrinology 149:4137–4141
Deleuze C, Duvoid A, Hussy N (1998) Properties and glial origin of osmotic-dependent release of
taurine from the rat supraoptic nucleus. J Physiol 507(Pt 2):463–471
Elmariah SB, Oh EJ, Hughes EG, Balice-Gordon RJ (2005) Astrocytes regulate inhibitory synapse
formation via Trk-mediated modulation of postsynaptic GABAA receptors. J Neurosci
25:3638–3650
Ero C, Gewaltig MO, Keller D, Markram H (2018) A cell atlas for the mouse brain. Front
Neuroinform 12:84
Fernandez-Galaz MC, Morschl E, Chowen JA, Torres-Aleman I, Naftolin F, Garcia-Segura LM
(1997) Role of astroglia and insulin-like growth factor-I in gonadal hormone-dependent synap-
tic plasticity. Brain Res Bull 44:525–531
Fuente-Martin E, Garcia-Caceres C, Granado M, de Ceballos ML, Sanchez-Garrido MA, Sarman B,
Liu ZW, Dietrich MO, Tena-Sempere M, Argente-Arizon P, Diaz F, Argente J, Horvath TL,
Chowen JA (2012) Leptin regulates glutamate and glucose transporters in hypothalamic
astrocytes. J Clin Invest 122:3900–3913
Gao Q, Mezei G, Nie Y, Rao Y, Choi CS, Bechmann I, Leranth C, Toran-Allerand D, Priest CA,
Roberts JL, Gao XB, Mobbs C, Shulman GI, Diano S, Horvath TL (2007) Anorectic estrogen
mimics leptin’s effect on the rewiring of melanocortin cells and Stat3 signaling in obese
animals. Nat Med 13:89–94
Garcia-Caceres C, Fuente-Martin E, Argente J, Chowen JA (2012) Emerging role of glial cells in
the control of body weight. Mol Metab 1:37–46
Garcia-Segura LM, McCarthy MM (2004) Minireview: Role of glia in neuroendocrine function.
Endocrinology 145:1082–1086
Garcia-Segura LM, Duenas M, Busiguina S, Naftolin F, Chowen JA (1995) Gonadal hormone
regulation of neuronal-glial interactions in the developing neuroendocrine hypothalamus. J
Steroid Biochem Mol Biol 53:293–298
Garcia-Segura LM, Lorenz B, DonCarlos LL (2008) The role of glia in the hypothalamus:
implications for gonadal steroid feedback and reproductive neuroendocrine output. Reproduc-
tion 135(4):419–429
Haan N, Goodman T, Najdi-Samiei A, Stratford CM, Rice R, El Agha E, Bellusci S, Hajihosseini
MK (2013) Fgf10-expressing tanycytes add new neurons to the appetite/energy-balance
regulating centers of the postnatal and adult hypothalamus. J Neurosci 33:6170–6180
388 L. L. DonCarlos and J. A. Chowen
Mong JA, Kurzweil RL, Davis AM, Rocca MS, McCarthy MM (1996) Evidence for sexual
differentiation of glia in rat brain. Horm Behav 30:553–562
Mong JA, Glaser E, McCarthy MM (1999) Gonadal steroids promote glial differentiation and alter
neuronal morphology in the developing hypothalamus in a regionally specific manner. J
Neurosci 19:1464–1472
Morel L, Higashimori H, Tolman M, Yang Y (2014) VGluT1+ neuronal glutamatergic signaling
regulates postnatal developmental maturation of cortical protoplasmic astroglia. J Neurosci
34:10950–10962
Morel L, Chiang MSR, Higashimori H, Shoneye T, Iyer LK, Yelick J, Tai A, Yang Y (2017)
Molecular and functional properties of regional astrocytes in the adult brain. J Neurosci
37:8706–8717
Morrell JI, Krieger MS, Pfaff DW (1986) Quantitative autoradiographic analysis of estradiol
retention by cells in the preoptic area, hypothalamus and amygdala. Exp Brain Res 62: 343–354
Morton GJ, Cummings DE, Baskin DG, Barsh GS, Schwartz MW (2006) Central nervous system
control of food intake and body weight. Nature 443:289–295
Muller A, Hauk TG, Fischer D (2007) Astrocyte-derived CNTF switches mature RGCs to a
regenerative state following inflammatory stimulation. Brain 130:3308–3320
Muller-Fielitz H, Stahr M, Bernau M, Richter M, Abele S, Krajka V, Benzin A, Wenzel J, Kalies K,
Mittag J, Heuer H, Offermanns S, Schwaninger M (2017) Tanycytes control the hormonal
output of the hypothalamic-pituitary-thyroid axis. Nat Commun 8:484
Munekawa K, Tamada Y, Iijima N, Hayashi S, Ishihara A, Inoue K, Tanaka M, Ibata Y (2000)
Development of astroglial elements in the suprachiasmatic nucleus of the rat: with special
reference to the involvement of the optic nerve. Exp Neurol 166:44–51
Naftolin F, Mor G, Horvath TL, Luquin S, Fajer AB, Kohen F, Garcia-Segura LM (1996) Synaptic
remodeling in the arcuate nucleus during the estrous cycle is induced by estrogen and precedes
the preovulatory gonadotropin surge. Endocrinology 137:5576–5580
Noctor SC, Flint AC, Weissman TA, Dammerman RS, Kriegstein AR (2001) Neurons derived from
radial glial cells establish radial units in neocortex. Nature 409:714–720
Oberheim NA, Takano T, Han X, He W, Lin JHC, Wang F, Qiwu Xu Q, Wyatt JD, Pilcher W,
Ojemann JG, Ransom BR, Goldman SA, Nedergaard M (2009) Uniquely hominid features of
adult human astrocytes. J Neurosci 29(10):3276
Ojeda SR (1994) The neurobiology of mammalian puberty: has the contribution of glial cells been
underestimated? J NIH Res 6:51–56
Ojeda SR, Prevot V, Heger S, Lomniczi A, Dziedzic B, Mungenast A (2003) Glia-to-neuron
signaling and the neuroendocrine control of female puberty. Ann Med 35:244–255
Ojeda SR, Lomniczi A, Sandau US (2008) Glial-gonadotrophin hormone (GnRH) neurone
interactions in the median eminence and the control of GnRH secretion. J Neuroendocrinol
20:732–742
Olmos G, Naftolin F, Perez J, Tranque PA, Garcia-Segura LM (1989) Synaptic remodeling in the
rat arcuate nucleus during the estrous cycle. Neuroscience 32:663–667
Parkash J, Messina A, Langlet F, Cimino I, Loyens A, Mazur D, Gallet S, Balland E, Malone SA,
Pralong F, Cagnoni G, Schellino R, De Marchis S, Mazzone M, Pasterkamp RJ, Tamagnone L,
Prevot V, Giacobini P (2015) Semaphorin7A regulates neuroglial plasticity in the adult hypo-
thalamic median eminence. Nat Commun 6:6385
Pelluru D, Konadhode RR, Bhat NR, Shiromani PJ (2016) Optogenetic stimulation of astrocytes in
the posterior hypothalamus increases sleep at night in C57BL/6J mice. Eur J Neurosci
43:1298–1306
Perea G, Navarrete M, Araque A (2009) Tripartite synapses: astrocytes process and control synaptic
information. Trends Neurosci 32:421–431
Perez J, Naftolin F, Garcia Segura LM (1990) Sexual differentiation of synaptic connectivity and
neuronal plasma membrane in the arcuate nucleus of the rat hypothalamus. Brain Res
527:116–122
390 L. L. DonCarlos and J. A. Chowen
Perlmutter LS, Tweedle CD, Hatton GI (1985) Neuronal/glial plasticity in the supraoptic dendritic
zone in response to acute and chronic dehydration. Brain Res 361:225–232
Pfrieger FW, Barres BA (1997) Synaptic efficacy enhanced by glial cells in vitro. Science
277:1684–1687
Pinto S, Roseberry AG, Liu H, Diano S, Shanabrough M, Cai X, Friedman JM, Horvath TL (2004)
Rapid rewiring of arcuate nucleus feeding circuits by leptin. Science 304(5667):110–115
Poluch S, Juliano SL (2007) A normal radial glial scaffold is necessary for migration of
interneurons during neocortical development. Glia 55:822–830
Prevot V, Croix D, Bouret S, Dutoit S, Tramu G, Stefano GB, Beauvillain JC (1999) Definitive
evidence for the existence of morphological plasticity in the external zone of the median
eminence during the rat estrous cycle: implication of neuro-glio-endothelial interactions in
gonadotropin-releasing hormone release. Neuroscience 94:809–819
Prevot V, Dehouck B, Sharif A, Ciofi P, Giacobini P, Clasadonte J (2018) The versatile tanycyte: a
hypothalamic integrator of reproduction and energy metabolism. Endocr Rev 39:333–368
Ramirez D, Saba J, Carniglia L, Durand D, Lasaga M, Caruso C (2015) Melanocortin 4 receptor
activates ERK-cFos pathway to increase brain-derived neurotrophic factor expression in rat
astrocytes and hypothalamus. Mol Cell Endocrinol 411:28–37
Robins SC, Stewart I, McNay DE, Taylor V, Giachino C, Goetz M, Ninkovic J, Briancon N,
Maratos-Flier E, Flier JS, Kokoeva MV, Placzek M (2013) alpha-Tanycytes of the adult
hypothalamic third ventricle include distinct populations of FGF-responsive neural progenitors.
Nat Commun 4:2049
Rodriguez EM, Blazquez JL, Pastor FE, Pelaez B, Pena P, Peruzzo B, Amat P (2005) Hypothalamic
tanycytes: a key component of brain-endocrine interaction. Int Rev Cytol 247:89–164
Rottkamp DM, Rudenko IA, Maier MT, Roshanbin S, Yulyaningsih E, Perez L, Valdearcos M,
Chua S, Koliwad SK, Xu AW (2015) Leptin potentiates astrogenesis in the developing
hypothalamus. Mol Metab 4:881–889
Silverman RC, Gibson MJ, Silverman AJ (1991) Relationship of glia to GnRH axonal outgrowth
from third ventricular grafts in hpg hosts. Exp Neurol 114:259–274
Sims RE, Butcher JB, Parri HR, Glazewski S (2015) Astrocyte and neuronal plasticity in the
somatosensory system. Neural Plast 2015:732014
Souttou S, Benabdesselam R, Siqueiros-Marquez L, Sifi M, Deliba M, Vacca O, Charles-Messance-
H, Vaillend C, Rendon A, Guillonneau X, Dorbani-Mamine L (2019) Expression and localiza-
tion of dystrophins and beta-dystroglycan in the hypothalamic supraoptic nuclei of rat from birth
to adulthood. Acta Histochem 121:218–226
Stern JE (2015) Neuroendocrine-autonomic integration in the paraventricular nucleus: novel roles
for dendritically released neuropeptides. J Neuroendocrinol 27:487–497
Suarez I, Fernandez B, Bodega G, Tranque P, Olmos G, Garcia-Segura LM (1987) Postnatal
development of glial fibrillary acidic protein immunoreactivity in the hamster arcuate nucleus.
Brain Res 465:89–95
Sweeney P, Qi Y, Xu Z, Yang Y (2016) Activation of hypothalamic astrocytes suppresses feeding
without altering emotional states. Glia 64:2263–2273
Theodosis DT, Poulain DA (1984) Evidence for structural plasticity in the supraoptic nucleus of the
rat hypothalamus in relation to gestation and lactation. Neuroscience 11:183–193
Theodosis DT, Poulain DA (1999) Contribution of astrocytes to activity-dependent structural
plasticity in the adult brain. Adv Exp Med Biol 468:175–182
Theodosis DT, Montagnese C, Rodriguez F, Vincent JD, Poulain DA (1986) Oxytocin induces
morphological plasticity in the adult hypothalamo-neurohypophysial system. Nature
322:738–740
Tweedle CD, Hatton GI (1977) Ultrastructural changes in rat hypothalamic neurosecretory cells and
their associated glia during minimal dehydration and rehydration. Cell Tissue Res 181:59–72
Verkhratsky A, Nedergaard M (2018) Physiology of astroglia. Physiol Rev 98:239–389
Wang YF, Hatton GI (2007) Interaction of extracellular signal-regulated protein kinase 1/2 with
actin cytoskeleton in supraoptic oxytocin neurons and astrocytes: role in burst firing. J Neurosci
27:13822–13834
14 Astrocytes and Development of Neuroendocrine Circuits 391
Wang YF, Hatton GI (2009) Astrocytic plasticity and patterned oxytocin neuronal activity: dynamic
interactions. J Neurosci 29:1743–1754
Witkin JW, Ferin M, Popilskis SJ, Silverman AJ (1991) Effects of gonadal steroids on the
ultrastructure of GnRH neurons in the rhesus monkey: synaptic input and glial apposition.
Endocrinology 129:1083–1092
Wolosker H, Balu DT, Coyle JT (2016) The rise and fall of the D-serine-mediated gliotransmission
hypothesis. Trends Neurosci 39:712–721
Wolosker H, Balu DT, Coyle JT (2017) Astroglial versus neuronal D-serine: check your controls!
Trends Neurosci 40:520–522
Wu H, Friedman WJ, Dreyfus CF (2004) Differential regulation of neurotrophin expression in basal
forebrain astrocytes by neuronal signals. J Neurosci Res 76:76–85
Wu YE, Pan L, Zuo Y, Li X, Hong W (2017) Detecting activated cell populations using single-cell
RNA-seq. Neuron 96:313–329.e316
Zamanian JL, Xu L, Foo LC, Nouri N, Zhou L, Giffard RG, Barres BA (2012) Genomic analysis of
reactive astrogliosis. J Neurosci 32(18):6391–6410
Zhang Y, Barres BA (2010) Astrocyte heterogeneity: an underappreciated topic in neurobiology.
Curr Opin Neurobiol 20:588–594
Zhang Y, Reichel JM, Han C, Zuniga-Hertz JP, Cai D (2017) Astrocytic process plasticity and
IKKbeta/NF-kappaB in central control of blood glucose, blood pressure, and body weight. Cell
Metab 25:1091–1102.e1094
Origins of Sex Differentiation of Brain
and Behavior 15
Margaret M. McCarthy
Abstract
Sex differences in adult brain and behavior are often established early in devel-
opment when the brain is remarkably immature. In adults sex differences take on
many forms including latent variables, dimorphisms, frequency, and more.
Androgens and estrogens from the developing male fetal testis masculinize the
brain, so that physiology and behavior are in synchrony and harmony with
gonadal phenotype. This differentiation process occurs during a critical window
in males, but females remain sensitive to exogenous steroid treatment for a longer
developmental period. The cellular mechanisms of sexual differentiation are
diverse but often involve inflammatory signaling molecules and immune cells.
This may have consequences for the higher vulnerability of males to neurological
and neuropsychiatric disorders of development.
Keywords
Amygdala · Androgens · Estrogens · Hypothalamus · Neuroimmune · Preoptic
area
15.1 Introduction
The brain is a reproductive organ, essential for both the control of gamete production
by the gonads and the physiology and behaviors required for achieving fertilization,
gestation, parturition, and parenting. By definition, these are different in males and
females since sex is defined by gamete size, production, and delivery. But determin-
ing how the brain drives the different physiologies and behaviors of males and
females is confounded by the inherent complexity of the nervous system and its wide
M. M. McCarthy (*)
Department of Pharmacology, University of Maryland School of Medicine, Baltimore, MD, USA
e-mail: [email protected]
The story of sex differences in the brain is one of on-again off-again. Unbeknownst
to many students of behavioral neuroendocrinology, in the 1960s and 1970s it was
considered radical to say there was any difference in the brains of males and females,
so radical that the first two studies in mammals were published in Science (Pfaff
1966; Raisman and Field 1971), and both found only very small and seemingly
insignificant differences, thereby confirming that male and female brains were
largely the same. This was prior to a full understanding of the neural control of
gonadotropin secretion from the pituitary, and indeed even the relationship between
the brain and the “master gland” was misunderstood until the 1950s (Melmed 2011).
In the 1960s, Charles Barraclough made the surprising observation that treating
newborn female rat pups with high doses of androgen resulted in a sterile adult
(Barraclough and Gorski 1961). He assumed this was the result of a direct effect on
the immature ovary, but through systematic investigation determined that it was
actually the result of dysregulation in the brain, and resulted from the loss of the
ability to release a surge of luteinizing hormone from the pituitary. This illustrated
for the first time that hormones early in development could “program” the brain
differently in males and females, but there were no clear neuroanatomical
underpinnings to the effect.
The story changed dramatically when studies on song birds revealed robust
differences in the nuclei in the brain that control song. This was intuited by the
fact that males sing complex melodies while females mostly tweet, but the fact that
this behavioral difference so clearly mapped onto neuroanatomy had not been
previously demonstrated (Nottebohm and Arnold 1976). The observation in birds
reopened investigation into mammals and several neuroanatomical sex differences
related to reproduction were subsequently identified. Among these are the sexually
dimorphic nucleus of the preoptic area (SDN-POA), the spinal nucleus of the
15 Origins of Sex Differentiation of Brain and Behavior 395
(continued)
396 M. M. McCarthy
deteriorating. Tracey Shors and colleagues determined this was correlated with a
change in the synaptic profile of hippocampal neurons known to be associated
with that particular form of learning (Shors 2006). An additional means by
which sex differences may be hidden is by appearing the same in presentation in
the two sexes but differing in prevalence. Alzheimer’s disease is the same
disease in men and women, but more frequently diagnosed in women, and not
just because they live longer. Conversely, stuttering is more common in boys
than girls. This type of sex difference has been referred to as a population sex
difference.
(c) Yet another form that sex differences take is in the existence of convergence,
also referred to as compensation and first articulated by Geert De Vries (2004).
In this instance, males and females are at a different baseline or set point, but
under the appropriate circumstance can converge on the same endpoint. Some-
times this occurs because of a lack of the necessary physiology in one sex or the
other. For instance, nurturing behavior toward young by adult males is not a
general feature of the behavioral repertoire of most rodent species, including
toward their own progeny, whereas females who have given birth universally
care for their offspring. Pregnancy and parturition involve marked physiological
and hormonal changes that are critical to inducing maternal behavior. In
mammals, males are deprived of this experience, both physiologically and
hence behaviorally as well, with one notable exception. Males in one particular
species of monogamous voles have evolved a unique vasopressinergic circuitry
that promotes paternal behavior, meaning they have compensated for the lack of
exposure to the hormones of pregnancy by generating a new independent
regulatory circuit. Catherine Woolley has articulated a related but distinct
principle that she also refers to as latent sex differences. A latent sex difference
is one that is not immediately apparent. For instance, measures of synapse
function of hippocampal pyramidal neurons are the same in males and females,
but the underlying cellular mechanisms mediating synaptic plasticity are highly
distinct, involving separate signal transduction pathways or different receptors
all together (Huang and Woolley 2012). It is essential to conceptualize the
various ways a sex difference can manifest when attempting to unravel and
understand the origins of either a developmental or adult response profile
(Fig. 15.1).
Fig. 15.1 Categories of sex differences. There are multiple ways in which the sexes can differ and
these begin with whether the dependent variable (endpoint) is the same or different. Left. When the
endpoint is different, it may be sexually dimorphic, meaning it takes on a different form in males
versus females, or it may be the average measure of the endpoint is significantly different in the two
sexes. A third possibility is that the two sexes are similar at baseline but diverge to different
endpoints upon a challenge or with aging. Right. When the endpoint is the same it can be the result
of different pathways or mechanisms, which are referred to as latent variables. Alternatively, the
sexes may be divergent at baseline but converge to the same endpoint following challenge or evolve
a mechanism to achieve the same endpoint via distinct pathway (therefore involving latent
variables). Lastly, the endpoint may be the same in males and females but occur more frequently
in the population in one sex versus the other
exclusion of another. It also usually means the decision point is restricted to a critical
period during which the relevant environmental stimulus is more salient and the
biological processes mediating fate determination are in synchrony. One of the best
examples is the coinciding events of eye opening and innervation of the lateral
geniculate in altricial mammals. Light is the essential external stimulus, and coordi-
nated neuronal activity is the essential internal biological response. The result is a
functioning visual system. If light is withheld during the critical period, the animal
will be forever blind. Going one step further, if an animal is raised during the critical
period in an environment consisting of only vertical lines, no horizontal lines, it will
be forever blind to horizontal lines. In this way, the nervous system fine tunes its
responses during the critical period, and those responses are then permanently
embedded in the brain.
In the case of the visual system, the environmental signal is external light. But
internal signals can also drive critical periods of developmental programming. The
amount and type of nutrition experienced prenatally programs future metabolomics.
Similarly, the amount and type of stress before and shortly after birth impact adult
behavioral and physiological phenotypes associated with anxiety, emotionality, and
activation of the hypothalamic–pituitary–adrenal axis. In both these instances, males
and females may differ in sensitivity or other aspects of critical period programming,
15 Origins of Sex Differentiation of Brain and Behavior 399
Fig. 15.2 Critical and sensitive periods for sexual differentiation in rodents and humans. The
critical period for sexual differentiation begins with the onset of testicular androgen production by
the developing fetus in rodents and humans. This is deemed a critical period because if androgens
levels do not rise at this time in males, the brain will fail to be masculinized. The sensitive period is
defined as that time during which a female can be treated with exogenous steroid (testosterone or
estrogen, or other masculinizing agents) and has the potential to be phenotypically converted to the
male for a particular endpoint. In rodents this extends into the first postnatal week but may vary by
endpoint. In humans, we do not know when or if there is an analogous sensitive period
but they both are exposed to the same salient internal or external stimuli. The case of
sexual differentiation of the brain is unique, however, in that only one sex is exposed
to the essential agent that drives the developmental program down a particular path.
That essential agent is testosterone from the testis of the fetal male (Fig. 15.2). Here
the goal is to differentiate male from female in order to assure that adult neurophysi-
ology and behavior is in synchrony with gonadal phenotype. Thus, males need to
continuously release small pulses of LH from the pituitary to maintain spermatogen-
esis and they need to seek out and mate with females and, in some species, compete
with other males and/or defend territories and vital resources. Females, by contrast,
must be capable of a massive surge in LH release in order to induce ovulation. They
must also seek out and mate with males, but usually preferably only one, and then
gestate, deliver, and nurture the offspring. Nurturing is a complex combination of
nesting, lactating, retrieving, grooming, and defending the young, each of which
involves distinct neural circuits and remain incompletely understood. This complex
divergence in physiology and behavior is programmed into the brain beginning prior
to birth of the individual, despite the actual endpoints not being manifest until
adulthood.
400 M. M. McCarthy
Up to this point, the process of masculinization has been emphasized because it is the
key “differentiation” event in sexual differentiation, meaning it is the process by
which the male is differentiated from the female. The feminization of both brain and
body is the developmental trajectory that occurs in the absence of testis and androgen
production, meaning an ovary is not required for female development (But it is
obviously essential to adult reproductive capacity). This does not mean that femini-
zation of the brain is not an active process, it surely is, but it is much harder to discern
what it is in the absence of some third “neither male nor female” phenotype. There is
some evidence of a later critical period in female brain development, during the
second postnatal week, that may involve estrogen production by the ovaries (Bakker
and Baum 2008), but unfortunately, this concept is not fully developed.
Perhaps the best angle on discerning what is feminization, is its polar opposite,
defeminization, which is an active process whereby the female phenotype is removed
(Fig. 15.3). This phenomenon is best illustrated, and perhaps limited to, the sex-typic
mating behavior seen in rats and mice where males show mounting of sexually
receptive females which respond with lordosis, a posture that allows the male to
intromit his penis. Since feminization is the default, the neural circuitry of lordosis
comes as preinstalled software. Removing that programming is achieved via defemi-
nization which is also driven by androgens aromatized to estrogens in developing
Fig. 15.3 Masculinization, feminization and defeminization. There are three distinct processes that
occur during sexual differentiation of the brain to regulate adult mating behavior. Upper. Mascu-
linization is induced by gonadal steroids of the fetal male testis to organize the neural substrate to be
responsive to androgens in adulthood and promote male mating behavior. Defeminization is also a
steroid induced process occurring during the same time as masculinization, and occurs via distinct
cellular mechanisms that result in erasing the innate program of feminization. As a result, sex
differences in adult mating behavior are maximized. Lower. Feminization occurs in the absence of
gonadal steroids and programs the neural substrate to respond to cyclical estrogens and progestins in
adulthood to promote female mating behavior. These events occur during distinct periods of
development, although the parameters surrounding a sensitive period for feminization remain
much less well defined than those for masculinization
15 Origins of Sex Differentiation of Brain and Behavior 401
males but via distinct cellular mechanisms (Schwarz and McCarthy 2008). Why such
a system has evolved is a mystery, and whether it applies outside the context of sex
behavior is debatable, but it tells us there are multiple independent processes occur-
ring simultaneously in the developing brain that assure as little overlap as possible
between males and females for certain key reproductive functions. Notably, there is
no parallel process of demasculinization in females, and when masculinization is
blocked in males it is best referred to as dysmasculinization, since it is the disruption
of a normal process, not itself a normal process.
Everyone is aware that puberty is a time of marked divergence for males and females
in terms of the hormonal milieu. But many are not as aware that as early as fetal life
there is a similar marked divergence created when the fetal testis becomes steroido-
genic and synthesizes sufficiently high levels of androgens that the steroid enters the
circulation and gains access to the fetal brain. In rodents, this process begins during
the last days of an approximately 3-week gestation and extends into the first hours
after birth (Weisz and Ward 1980). In human males, fetal steroidogenesis starts early
in the second trimester and is waning by birth but followed by a later active period
postnatally that is sometimes referred to as “mini-puberty” (Kuiri-Hanninen et al.
2014). The opening of the critical period for sexual differentiation of the brain is
operationally defined by the onset of testicular androgen production. Artificially
elevating androgens prior to the naturally occurring period of steroidogenesis is
largely without effect, as the essential biological processes are not yet in play. The
closing of the critical period is another matter. In males, the critical period closes
when the surge of androgen production ends. Once the system is fully exposed to
testosterone the programming process is initiated and cannot be undone, meaning
that within one day of birth masculinization is largely complete and the critical
period closed. Females, however, remain sensitive to the programming potential of
testosterone well into the postnatal period, up to a week (Fig. 15.2). If injected with
either testosterone or estradiol, the masculinization process is initiated, reflecting a
sensitive period as opposed to a critical period (McCarthy et al. 2018). This
seemingly lucky accident of timing in rodents provides an excellent tool for studying
the mechanisms of steroid-mediated programming, by essentially using the female
as a platform for exogenously induced masculinization. In nonhuman primates,
however, the end of the critical and the sensitive period are both prenatal (Wallen
and Baum 2002), confounding our ability to replicate mechanistic findings in
rodents and instead forcing us to consider other routes for translational insight,
which will be discussed below. Moreover, since we cannot experiment on humans,
we do not know if there is a postnatal sensitive period in addition to the prenatal
critical period.
The rodent model has many advantages, but there are confounds just as there are
in any “model.” The ability to induce the masculinization process in newborn
females is an important strength but since the naturally occurring process in males
402 M. M. McCarthy
begins in utero, the experimental females are in fact older when treatment occurs. At
this stage of development, even a few days is a lot of time for changes to occur. It is
possible to treat pregnant dams with steroids and thereby begin the exposure of
females in utero, but the endocrine milieu of late pregnancy is both robust and
dynamic in preparation for birth. Introducing exogenous steroids against this back-
drop is unavoidably confounded. In early studies, beginning with the iconic Phoenix
et al. (1959) study, androgen treatment of pregnant guinea pigs not only
masculinized the sexual behavior of female offspring but it also masculinized the
genitalia, an obvious confound that needs to be avoided if one wishes to understand
the masculinization of the brain in isolation from the body. Lastly, rodents give birth
to litters, meaning multiple offspring develop in two uterine horns where they are
packed in like peas in a pod. Evidence of just how sensitive the developing brain is to
gonadal steroids is found in the intrauterine position (IUP) phenomenon whereby
females who develop between two males are slightly masculinized and males who
develop between two females are slightly less masculinized relative to littermates
that sit between mixed-sex or same sex pups. The shift in masculinization is due to
androgen production of fetal males being so profuse that some is transferred to
nearest neighbors. This is a naturally occurring source of individual variation in
phenotype that is independent of genetics and is often overlooked. Endpoints that are
impacted by IUP include adult hormonal levels, aggression, sexual behavior, and
susceptibility to endocrine disruption. Limited evidence in humans suggests that
females from mixed-sex twins exhibit some evidence of masculinization (Ryan and
Vandenbergh 2002).
The famous geneticist Dobszhansky is often quoted for saying “Nothing in biology
makes sense except in the light of evolution.” Equally true would be “Nothing in
neuroscience makes sense except in the light of development.” The brain is the
slowest developing organ of the body and is designed to be impacted by the local
environment and experience as it matures. A fundamental goal of neuroscience is to
understand how the brain controls complex adult behaviors. We want to know how
the brain determines that now is the time to feed, now to run, now to mate, now to
flee, now to fight, and now to sleep. All of these decisions are made against a
backdrop of current context and past experience, but they are also impacted by how
the brain was wired from the beginning, during development. Development is a
simple word for a dynamic multifactorial robust process that is also sensitive to
perturbations, sometimes only during a specific window or time period. More
importantly, in development what happens first affects what happens next, which
affects what happens next and so on in an increasingly constrained one-way trajec-
tory. Development cannot be undone. It is the pervasive underpinning of all that
follows.
Sexual differentiation of the brain is a developmental process but one of the more
astonishing things about it is the timing, meaning how early. In humans our best
15 Origins of Sex Differentiation of Brain and Behavior 403
estimate of the critical period for sexual differentiation is early second trimester to
mid- to late-third trimester. This means the brain is being organized for behaviors
that will not be expressed for a decade or more, with all the intervening maturation of
this most complex organ. In rodents the critical period is more clearly defined due to
our ability to experiment, beginning with the onset of testicular hormones during the
last days of gestation and lasting until the first few days of life. But here too the brain
is remarkably immature during the process of differentiation. So how is it that these
early organizational events endure until adulthood?
The conceptualization of sexual differentiation has evolved along with our
expanding views of the brain and its capabilities. In the 1970s and 1980s the mature
brain was still considered essentially immutable, with no capacity for neurogenesis,
regeneration, or even new synapse formation. The brain was often interpreted as
analogous to a computer, being hard wired and programmed. Consistent with that, the
impact of sexual differentiation was referred to as building the “neuroarchitecture,”
establishing a “blue print” and “constructing” a circuit. But we now know the brain is
highly mutable, even plastic, with new neurons being born well into adulthood and
synapses coming and going with astonishing frequency. At first this view makes it
even more challenging to fathom how a developmental process in a brain that is
essentially a gelatinous mass can be programmed to regulate later complex social
behaviors. Here a second revolution in our thinking about the brain comes into play,
the role of epigenetic modifications.
15.8 Neuroepigenetics
Epigenetics literally means “above the genome” and thus refers to modifications to
the genetic code that impact gene expression. The dominant forms of epigenetic
modifications are changes to the histones that are integral to the chromatin around
the DNA and the addition of methyl groups to the nucleotides themselves. There are
a variety of changes to the histones but the most prevalent is the addition or removal
of acetyl and/or methyl groups to lysine residues. Changes to the DNA itself are
largely restricted to the addition of methyl groups and this occurs mostly but not
exclusively on cytosines proximal to guanines, so-called CpGs. The notion of
epigenetics has been with us for a long time, dating back to the geneticist
Waddington in the 1940s who invoked the term to explain how the same DNA in
every cell could lead to divergent cell states. There needed to be a mechanism by
which some genes would be permanently repressed while others were expressed.
This is required in order for a hepatocyte to remain a hepatocyte while an osteoclast
stays an osteoclast.
Epigenetics was rediscovered in the nervous system as a means to explain how
behavioral phenotypic variation could be transferred from dam to pup. More specifi-
cally, Michael Meaney and colleagues determined that rat dams that were highly
attentive to their pups produced adult females that were also highly attentive to their
pups. That this was not genetic in the classic sense was proven by cross fostering
pups born to low attentive dams to highly attentive dams and observing that they
404 M. M. McCarthy
briefly prevents further methylation during the first few postnatal days and then
raised to adulthood. In order to determine if their behavior was masculinized, the
females were ovariectomized and implanted with silastic capsules that release
testosterone. Once the circulating levels of testosterone in these females were on
par with that normally seen in males, they were tested for mating behavior in a
controlled setting using test females who were sexually receptive.
Intriguingly, these females were fully masculinized, behaving just as normal
males do, and examination of their brains at completion of the experiment revealed
the synaptic pattern was the same as that seen in males as well. Collectively these
data suggest that epigenetic modifications produced by steroid-mediated DNMT
activity during the critical period help to maintain a dendritic architecture that is
conducive to male-typic mating behavior in adulthood (Fig. 15.4). Of particular
interest, this is a steroid driven process but does not involve interaction of the steroid
and its receptor with the DNA that is being modified, it is more of an epigenetic
modification. The mechanism by which estradiol modulates DNMT activity is
unknown but there is no change in either the mRNA or the protein levels of
DNMT1, DNMT3, and DNMT3a. Steroids never cease to amaze.
Critical periods have a beginning and an end. The beginning is usually associated
with the onset of some stimulus, such as in the case of sexual differentiation, the
onset of steroidogenesis by the fetal testis. But the ends of critical periods are much
more elusive, usually being defined as that developmental time point at which the
brain no long responds to the requisite stimulus. But what determines the loss of
responsiveness? The observation of a change in DNA methylation offered a clue. If
hormone-responsive genes are epigenetically silenced at a particular developmental
time point, that could close the window of the critical/sensitive period. Indeed, this
does appear to be the case since in the aforementioned study. If females were treated
with the DNA demethylating drug after the normal close of the critical/sensitive
period, they were still masculinized as evidenced by adult brain and behavior.
Importantly, if females were treated with a masculinizing dose of steroid at the
same time point, there was no effect, indicating a loss of steroid responsiveness due
to DNA methylation. Thus, the close of the critical/sensitive period for sexual
differentiation may be mediated by epigenetic repression of hormone-responsive
genes that promote masculinization (Nugent et al. 2015).
In addition to the epigenetic effects of steroids described above, there are
examples of direct gene transcription mediated sexual differentiation. The amygdala
is a brain region critical to social behavior in general and a mediator of sex
differences in juvenile play behavior in particular. In males the cells of the amygdala
express more of the neuropeptide vasopressin, a sex difference that is programmed
developmentally by androgens. If a particular DNA methyl-binding protein called
MeCP2 is reduced during the critical period in males, the adult vasopressin level is
also reduced, indicating a neonatal epigenetic programming of adult gene expression
(Forbes-Lorman et al. 2012).
Given the central role of steroid receptors in both developmental and adult
sensitivity to gonadal hormones, it is reasonable to also ask if the genes coding for
estrogen receptors alpha and beta or progesterone receptor are also epigenetically
406 M. M. McCarthy
Fig. 15.4 Epigenetic programming of masculinization. Upper, males. The process of masculini-
zation of brain and behavior occurs weeks, months to decades prior to the expression of the
phenotype. One mechanism by which this early life programming is maintained is via epigenetic
modifications of the DNA in which methyl groups are added to cytosine nucleotides or via changes
to the surrounding chromatin, both of which impart long-term regulatory functions on gene
expression. In the developing POA elevated gonadal steroids in males inhibit the DNMT enzymes
that transfer a methyl group to cytosines, resulting in reduced DNA methylation compared to
females. Adult male mounting behavior is illustrated. Lower, females. The higher level of methyla-
tion of the female genome in the POA inhibits the gene expression profile essential for adult
expression of male sex behavior and may also provide the means by which the sensitive period is
closed. Adult female lordosis behavior and maternal behavior are illustrated
modified. There are studies of the promoter regions of all these receptors and there
are sex differences, but there is little agreement about the direction and magnitude of
the amount of methylation or, even in some cases, precisely which CpGs are
differentially methylated (McCarthy et al. 2017). As more studies are conducted, it
is becoming increasingly clear that the epigenetics of the brain truly are highly
malleable and that a sex difference can exist at one point in time in one brain region,
disappear and reappear at another time at different CpG’s within the same gene.
Moreover, some hormonally induced epigenetic changes do not appear until long
after hormone exposure. In what might be called an “epigenetic echo,” two indepen-
dent studies have found hormonally induced changes in DNA methylation that were
not apparent within days of steroid treatment in neonates but did appear in juveniles
or adults (Ghahramani et al. 2014; Nugent et al. 2010).
15 Origins of Sex Differentiation of Brain and Behavior 407
How this happens at a mechanistic level remains entirely unknown but highlights
one of the biggest challenges in our quest to understand epigenetic underpinnings of
sexual differentiation—our inability to repeatedly monitor changes in the same
animal. Many things that we seek to measure as neuroendocrinologists can be
measured repeatedly, for example, circulating corticosterone, or LH release, body
weight, even neuronal activity via indwelling electrodes. But epigenetic
modifications to the DNA and surrounding chromatin can only be assessed ex vivo.
Moreover, most techniques involve a mixture of cell types and thereby contribute
noise to the output and uncertainty to the interpretation. Lastly, the approaches with
the highest fidelity, such as deep sequencing, are not as widely available for most
laboratories as other techniques. Fortunately, all of these problems are improving in
the right direction and so while neuroepigenetics is still at the frontier of neurosci-
ence, it is a worthy frontier to explore.
Steroids drive masculinization of the developing brain, but how? Many investigators
have explored the mechanistic basis of sexual differentiation for many years, but it is
fair to say that no clear unifying theory emerged from the effort and a general
conclusion of “it’s complicated” pervaded the field. The majority of work focused on
neurotransmitter systems for the obvious reason that they are central to the control of
behavior. Disruption of a particular neurotransmitter system will often impair normal
masculinization, but there was no documented incidence of being able to induce
masculinization in the absence of steroids. This stalemate was ended with the
discovery that the key masculinizing agent in at least one case—the dendritic
synaptic profile of the POA and adult mating behavior—was not a neurotransmitter
at all, but instead the membrane derived inflammatory signaling molecule, prosta-
glandin E2, also called PGE2 (Amateau and McCarthy 2004). Prostaglandins are
synthesized from arachidonic acid in the cell membrane by the cyclooxygenase
enzymes, COX1 and COX2. Males have higher levels of both enzymes in the
POA, leading to higher PGE2 production, which binds to the EP2 and EP4 receptors,
G-protein coupled receptors that activate adenylate cyclase and cAMP production.
Subsequent activation of protein kinase A (PKA) and phosphorylation of key
subunits of glutamate receptors induces clustering of glutamate receptors at the
membrane and the formation and stabilization of dendritic spines (Lenz et al.
2011). Many cells throughout the body are capable of synthesizing prostaglandins,
but of particular note are microglia, the brain’s innate immune system. These
macrophage-lineage cells are resident in the brain and they tile throughout, meaning
they spread themselves out so that each cell occupies a limited domain that does not
overlap with other microglia but the entire brain is in contact with at least one
microglial cell. They also respond to and produce prostaglandins, and, do so more in
males than females when located within the POA. Indeed, the microglia of the POA
are essential for normal masculinization to occur. If they are either depleted or
408 M. M. McCarthy
inhibited, the male brain develops without the capacity for male sexual behavior
(Lenz et al. 2013; VanRyzin et al. 2018).
The discovery that an innate immune cell is essential for normal masculinization
of brain and behavior was a surprise. But this surprise was surpassed with the
additional discovery that another innate immune cell, the mast cell, which is derived
from precursors made in the bone marrow, is rather the true mastermind behind the
immune-driven process of masculinization (Lenz et al. 2018). Mast cells are not
ubiquitous throughout the brain, instead being clustered in the meninges, and located
deep in the neuropil of the POA. Importantly, there are more in the male POA than in
the female. Remarkably, these cells number only in the tens, but when stimulated to
release histamine they signal to neighboring microglia which then increase produc-
tion of prostaglandins and the entire cascade leading to masculinization of brain and
behavior is initiated (Fig. 15.5). If the mast cells are prevented from releasing
histamine, the train is stopped and the brain dysmasculinized. Blocking the degran-
ulation of mast cells requires pharmacological treatment. Evidence for a more
nuanced role for mast cells came from inducing an allergic reaction in pregnant
dams and observing that the mast cells of the POA in the offspring were more
activated. Moreover, in adulthood the females were shifted towards a masculinized
behavioral profile while the male offspring were less masculinized (Lenz et al.
2019). In both cases the behavior of the males and females was within the normal
range, if that range includes both sexes, suggesting this is a potential source of
natural variation in behavior. In summary, there are multiple cell types involved in
the process of masculinizing the POA, and at least two of them are non-neuronal
cells; microglia and mast cells.
Microglia have proven to be equally important to another sexually differentiated
behavior, juvenile social play, but involving a completely different mechanism. Play
in rodents, sometimes called “rough-and-tumble” play is a complex social behavior
that may involve dyads of the same or different sex, or could involve large groups of
six or more animals (Argue and McCarthy 2015). The medial amygdala has long
been known as a key brain region in the neural circuitry of play and appears to be the
critical node for driving a sex difference in which males play more intensely and
frequently than females. How this occurs, however, was unknown until recently
when it was discovered that the number of astrocytes in the medial amygdala is
tightly controlled and is different between males and females, there being more in
females. This is not because more astrocytes are born in the female amygdala,
however, instead it is because the microglia of this brain region in males phagocy-
tose recently born astrocytes, thereby eliminating a portion of the population before
it has time to mature (VanRyzin et al. 2019). In this instance, there is no role for
prostaglandins or histamine but instead another class of membrane derived signaling
molecules, the endocannabinoids. These short-lived local signaling molecules are
found at a higher concentration in the developing amygdala of males and promote
phagocytosis by microglia. The end result is a neural circuitry that is more conducive
to rough-and-tumble play in males.
15 Origins of Sex Differentiation of Brain and Behavior 409
Fig. 15.5 Inflammatory signaling molecules and masculinization. In the preoptic area (POA),
sexual differentiation of synaptic density begins with hormonal regulation (by testosterone, T, or
estradiol, E2) of gene expression for the cyclooxygenase enzymes (COX-2) which make prosta-
glandin E2 (PGE2), an inflammatory signaling molecule. However additional PGE2 from the
microglia (brown cells), the brain’s own immune system, is also required for the full masculiniza-
tion process. The production of PGE2 by microglia is initiated by histamine release from resident
mast cells which are more numerous in males and thereby drive higher PGE2 production which
induces glutamate release from neighboring astrocytes. The glutamate binds to AMPA receptors
which have been induced to cluster at the membrane surface by the actions of protein kinase A
which is activated by PGE2 binding to G-protein coupled receptors on neurons. The activation of
the AMPA receptors causes the formation of dendritic spine synapses. Thus at least four cell types
are involved in the masculinization of the preoptic area, and two of them are of non-neuronal origin
but instead are components of the immune system
15.10 Perspectives
Sex differences are established during the process of sexual differentiation which
occurs weeks, months and in some cases years before the physiology and behavior
are manifest. Epigenetic programming both establishes and maintains sex
differences but also provides for reversibility across the life span. Sex differences
come in a variety of forms and understanding the nature of any particular one is
essential to proper interpretation and experimental design. The mechanisms of
sexual differentiation are diverse but often include inflammatory signaling molecules
and immune cells. The discovery that the immune profile of portions of the
410 M. M. McCarthy
developing male brain is elevated, or, more activated, compared to the female, is still
highly novel but is leading to new vistas of exploration. Re-analyses of published
transcriptomic data in human fetal cortex implicated greater expression of inflam-
matory pathways in males, and higher still in adult autistic males (Werling et al.
2016). This has led to speculation of a causal connection between the elevated
inflammatory state of the developing male brain and the higher risk for
neurodevelopmental psychiatric disorders such as autism spectrum disorders,
schizophrenia and attention deficit and hyperactivity disorders (McCarthy 2018).
Importantly, the researchers conducting the transcriptomic analysis in humans were
greatly aided in their ability to interpret their findings as genuine in light of the basic
science research done in animals. This is an important way in which so-called “blue
sky” research can have an impact.
We have likely seen just the tip of the proverbial iceberg of sex differences in
adult brain and behavior that have origins in development. Identifying the cellular
and molecular means by which early differences are established is essential to
understanding how they manifest in adulthood. Attention to developmental pro-
cesses as they are occurring is essential to elucidating the mechanisms by which sex
differences are established and ultimately maintained.
Key References
Amateau and McCarthy (2004)—This was the first report in which masculinization
of a female brain was achieved by a downstream signaling molecule from steroid
action, as opposed to treating directly with steroids. This opened a new era in
understanding of the novel mechanisms by which steroids can impact the brain,
including through the enhancement of immune system signaling.
Beery and Zucker (2010)—An in depth analyses of a large number of publications
across a wide swatch of biomedical research revealed the over reliance on male
animals, especially in neuroscience. Almost six times as many studies used
exclusively male rats and mice. Subsequent to this publication the NIH mandated
the incorporation of sex as a biological variable in all preclinical research.
De Vries (2004)—This minireview proposed the novel idea that sometimes the two
sexes evolve mechanisms to try and be more similar, not more different. Using
the example of the biparental vole, DeVries highlights how males have evolved a
new neural circuit that promotes parenting and is active in the absence of the
hormones of pregnancy and parturition that are so important to females.
Nottebohm and Arnold (1976)—Arguably the first robust neuroanatomical sex
difference discovered in a vertebrate brain, this land mark publication led to
increased scrutiny of the mammalian brain and the discovery of multiple sex
differences in the size of various brain regions and subnuclei.
Nugent et al. (2015)—One of a small number of studies that revealed sex differences
in the epigenome and in this case demonstrates functional significance for the
sexual differentiation of sexual behavior.
15 Origins of Sex Differentiation of Brain and Behavior 411
References
Amateau SK, McCarthy MM (2004) Induction of PGE(2) by estradiol mediates developmental
masculinization of sex behavior. Nat Neurosci 7(6):643–650
Argue KJ, McCarthy MM (2015) Characterization of juvenile play in rats: importance of sex of self
and sex of partner. Biol Sex Differ 6:16
Bakker J, Baum MJ (2008) Role for estradiol in female-typical brain and behavioral sexual
differentiation. Front Neuroendocrinol 29(1):1–16. https://doi.org/10.1016/j.yfrne.2007.06.001
Barraclough CA, Gorski RA (1961) Evidence that the hypothalamus is responsible for androgen-
induced sterility in the female rat. Endocrinology 68:68–79
Beery AK, Zucker I (2010) Sex bias in neuroscience and biomedical research. Neurosci Biobehav
Rev 35:565–572. https://doi.org/10.1016/j.neubiorev.2010.07.002
Clayton JA, Collins FS (2014) Policy: NIH to balance sex in cell and animal studies. Nature 509
(7500):282–283
De Vries GJ (2004) Minireview: sex differences in adult and developing brains: compensation,
compensation, compensation. Endocrinology 145(3):1063–1068
Forbes-Lorman RM, Rautio JJ, Kurian JR, Auger AP, Auger CJ (2012) Neonatal MeCP2 is
important for the organization of sex differences in vasopressin expression. Epigenetics 7
(3):230–238. https://doi.org/10.4161/epi.7.3.19265
Ghahramani NM, Ngun TC, Chen PY, Tian Y, Krishnan S, Muir S et al (2014) The effects of
perinatal testosterone exposure on the DNA methylome of the mouse brain are late-emerging.
Biol Sex Differ 5:8. https://doi.org/10.1186/2042-6410-5-8
Gould E, Woolley CS, Frankfurt M, McEwen BS (1990) Gonadal steroids regulate dendritic spine
density in hippocampal pyramidal cells in adulthood. Neuroscience 10:1286–1291
Huang GZ, Woolley CS (2012) Estradiol acutely suppresses inhibition in the hippocampus through
a sex-specific endocannabinoid and mGluR-dependent mechanism. Neuron 74(5):801–808.
https://doi.org/10.1016/j.neuron.2012.03.035
Hull EM, Dominguez JM (2007) Sexual behavior in male rodents. Horm Behav 52(1):45–55.
https://doi.org/10.1016/j.yhbeh.2007.03.030
Kuiri-Hanninen T, Sankilampi U, Dunkel L (2014) Activation of the hypothalamic-pituitary-
gonadal axis in infancy: minipuberty. Horm Res Paediatr 82(2):73–80. https://doi.org/10.
1159/000362414
Lenz KM, Wright CL, Martin RC, McCarthy MM (2011) Prostaglandin E regulates AMPA
receptor phosphorylation and promotes membrane insertion in preoptic area neurons and glia
during sexual differentiation. PLoS One 6(4):e18500. https://doi.org/10.1371/journal.pone.
0018500
Lenz KM, Nugent BM, Haliyur R, McCarthy MM (2013) Microglia are essential to masculinization
of brain and behavior. J Neurosci 33(7):2761–2772
Lenz KM, Pickett LA, Wright CL, Davis KT, Joshi A, McCarthy MM (2018) Mast cells in the
developing brain determine adult sexual behavior. J Neurosci 38(37):8044–8059. https://doi.
org/10.1523/JNEUROSCI.1176-18.2018
Lenz KM, Pickett LA, Wright CL, Galan A, McCarthy MM (2019) Prenatal allergen exposure
perturbs sexual differentiation and programs lifelong changes in adult social and sexual behav-
ior. Sci Rep 9:4837
412 M. M. McCarthy
Lomniczi A, Loche A, Castellano JM, Ronnekleiv OK, Bosch M, Kaidar G et al (2013) Epigenetic
control of female puberty. Nat Neurosci 16(3):281–289. https://doi.org/10.1038/nn.3319
McCarthy MM (2018) Sex differences in neuroimmunity as an inherent risk factor.
Neuropsychopharmacology 44(1):38-44. https://doi.org/10.1038/s41386-018-0138-1
McCarthy MM, Rissman EF (2014) Epigenetics of reproduction. In: Plant T, Zeleznik AJ (eds)
Knobil & Neill’s physiology of reproduction, vol 2, 4th edn. Academic, London, pp 2439–2501
McCarthy MM, Nugent BM, Lenz KM (2017) Neuroimmunology and neuroepigenetics in the
establishment of sex differences in the brain. Nat Rev Neurosci 18(8):471–484. https://doi.org/
10.1038/nrn.2017.61
McCarthy MM, Herold K, Stockman SL (2018) Fast, furious and enduring: sensitive versus critical
periods in sexual differentiation of the brain. Physiol Behav 187:13–19. https://doi.org/10.1016/
j.physbeh.2017.10.030
Melmed S (2011) The pituitary, 3rd edn. Academic, London
Nottebohm F, Arnold AP (1976) Sexual dimorphism in vocal control areas of the songbird brain.
Science 194:211–213
Nugent BM, Schwarz JM, McCarthy MM (2010) Hormonally mediated epigenetic changes to
steroid receptors in the developing brain: implications for sexual differentiation. Horm Behav.
https://doi.org/10.1016/j.yhbeh.2010.08.009
Nugent BM, Wright CL, Shetty AC, Hodes GE, Lenz KM, Mahurkar A et al (2015) Brain
feminization requires active repression of masculinization via DNA methylation. Nat Neurosci
18(5):690–697. https://doi.org/10.1038/nn.3988
Pfaff DW (1966) Morphological changes in the brains of adult male rats after neonatal castration.
J Endocrinol 36:415–416
Phoenix CH, Goy RW, Gerall AA, Young WC (1959) Organizing action of prenatally administered
testosterone proprionate on the tissues mediating mating behavior in the female guinea pig.
Endocrinol 65:369–382
Raisman G, Field PM (1971) Sexual dimorphism in the preoptic area of the rat. Science
173:731–733
Ryan BC, Vandenbergh JG (2002) Intrauterine position effects. Neurosci Biobehav Rev 26
(6):665–678
Schwarz JM, McCarthy MM (2008) Steroid-induced sexual differentiation of the brain: multiple
pathways, one goal. J Neurochem 105:1561–1572
Shors TJ (2006) Stressful experience and learning across the lifespan. Annu Rev Psychol 57:55–85
Smith CL, O’Malley BW (2004) Coregulator function: a key to understanding tissue specificity of
selective receptor modulators. Endocr Rev 25(1):45–71
Sweatt JD (2013) The emerging field of neuroepigenetics. Neuron 80(3):624–632. https://doi.org/
10.1016/j.neuron.2013.10.023
Szyf M, Weaver I, Meaney M (2007) Maternal care, the epigenome and phenotypic differences in
behavior. Reprod Toxicol 24(1):9–19. https://doi.org/10.1016/j.reprotox.2007.05.001
VanRyzin JW, Pickett LA, McCarthy MM (2018) Microglia: driving critical periods and sexual
differentiation of the brain. Dev Neurobiol 78(6):580–592. https://doi.org/10.1002/dneu.22569
VanRyzin JW, Marquardt AE, Argue KJ, Vecchiarelli HA, Ashton SE, Arambula SE, Hill MN,
McCarthy MM (2019) Microglial phagocytosis of newborn cells is induced by
endocannabinoids and sculpts sex differences in juvenile rat social play. Neuron 102:435–449
Wallen K, Baum MJ (2002) Masculinization and defeminization in altricial and precocial
mammals: comparative aspects of steroid hormone action. In: Pfaff D (ed) Hormones brain
and behavior, vol 5. Academic, London, pp 385–424
Weisz J, Ward IL (1980) Plasma testosterone and progesterone titers of pregnant rats, their male and
female fetuses and neonatal offspring. Endocrinology 106:306–313
Werling DM, Parikshak NN, Geschwind DH (2016) Gene expression in human brain implicates
sexually dimorphic pathways in autism spectrum disorders. Nat Commun 7:10717. https://doi.
org/10.1038/ncomms10717
Wright CL, Burks SR, McCarthy MM (2008) Identification of prostaglandin E2 receptors
mediating perinatal masculinization of adult sex behavior and neuroanatomical correlates.
Dev Neurobiol 68:1406–1419
Development and Modulation of Female
Reproductive Function by Circadian Signals 16
Neta Gotlieb, Jacob Moeller, and Lance J. Kriegsfeld
Abstract
Female reproductive success requires multifaceted, temporally coordinated hor-
mone secretion. The circadian timing system is central to this complex neuroen-
docrine regulation, with circadian disruption associated with irregular ovulatory
cycles, reduced fertility, increased miscarriage rates, and anomalous fetal devel-
opment. Because living in the modern world is associated with relatively chronic
circadian disruption due to limited sunlight exposure during the day and exposure
to artificial light at night, this issue is of broad concern. The master mammalian
circadian pacemaker in the suprachiasmatic nucleus communicates monosynap-
tically to neuroendocrine cells in the brain to mediate the hormonal events
required for ovulation and pregnancy maintenance. Likewise, maternal hormones
cross the placenta to drive rhythmic fetal processes critical for typical develop-
ment. The present overview describes the means by which the circadian timing
system integrates with the reproductive axis to regulate female reproductive
functioning. Likewise, the negative consequences of circadian disruptions on
female reproductive health, and the mechanisms underlying these deleterious
outcomes, are considered.
N. Gotlieb
Department of Psychology, University of California, Berkeley, Berkeley, CA, USA
J. Moeller
Graduate Group in Endocrinology, University of California, Berkeley, Berkeley, CA, USA
L. J. Kriegsfeld (*)
Department of Psychology, University of California, Berkeley, Berkeley, CA, USA
Department of Integrative Biology, University of California, Berkeley,, Berkeley, CA, USA
Graduate Group in Endocrinology, University of California, Berkeley, Berkeley, CA, USA
The Helen Wills Neuroscience Institute, University of California, Berkeley, Berkeley, CA, USA
e-mail: [email protected]
Keywords
Biological rhythms · Circadian · Hormone · Reproduction
16.1 Introduction
2008; Schoeller et al. 2016). Herein, we consider all phases of the female reproduc-
tive cycle, including ovulation, mating, pregnancy and fetal development, and
parturition and describe how the circadian system integrates with the reproductive
axis to mediate reproductive success. Likewise, we consider development in relevant
systems and circuits where applicable and the negative consequences of circadian
disruption during each stage of reproduction.
The circadian system consists of a master brain clock in the suprachiasmatic nucleus
(SCN) of the anterior hypothalamus that is synchronized to environmental time via a
direct retinal pathway. As indicated previously, although circadian rhythms are
endogenously generated, to be adaptive for an organism and anticipate environmen-
tal changes across the day, these endogenous cycles are entrained to environmental
time. Light entrains the SCN via retinal communication from rods and cones that, in
turn, target specialized, intrinsically photosensitive retinal ganglion cells containing
the photopigment, melanopsin (Hattar et al. 2002; Berson et al. 2002; Lucas et al.
2003; Panda et al. 2002). Successively, the SCN uses neural, diffusible, and auto-
nomic communication to convey timing information to the whole organism (Butler
et al. 2010; Buijs et al. 2013).
Circadian rhythms are a cell-autonomous property, with circadian rhythms being
generated by an autoregulatory transcription–translation feedback loop consisting of
clock genes and their protein products (Takahashi 2015; Honma 2018). The core
feedback loop begins in the morning with the clock protein, CLOCK, binding to
BMAL1 to drive the transcription of the Period (Per1 and Per2) and Cryptochrome
(Cry1 and Cry2) genes. Over the course of the day, Per and Cry transcripts are
translated into their respective proteins that inevitably feed back to the cell nucleus to
repress CLOCK: BMAL1-mediated transcription until the next morning when
transcription resumes. Whereas the Clock gene is constitutively expressed, an
additional feedback loop driven by the CLOCK: BMAL1 complex regulates
Bmal1 transcription through repression by Rev-erb/ and transcriptional activation
via retinoic acid receptors (RORs) (Fig. 16.1). Circadian timekeeping is a ubiquitous
property of cells throughout the brain and body, with virtually all cells exhibiting
circadian timekeeping (Honma 2018). Even when the SCN is isolated in culture, the
master clock maintains indefinite circadian rhythms at the tissue level due to unique
coupling among independent oscillators. In contrast, in the absence of master clock
communication or other entraining stimuli, extra-SCN brain loci and peripheral
organs exhibit loss of rhythmicity after several cycles (Yamazaki et al. 2000; Abe
et al. 2002). This loss of rhythmicity in extra-SCN systems results from loss of
coupling among cellular oscillators having slightly different periods (Welsh et al.
2004).
416 N. Gotlieb et al.
Fig. 16.1 The molecular clockwork. A simplified model of the intracellular mechanisms respon-
sible for mammalian circadian rhythm generation. The process begins when CLOCK and BMAL1
proteins dimerize to drive the transcription of the Per (Per1 and Per2) and Cry (Cry1 and Cry2)
genes. Throughout the day, PER and CRY proteins rise within the cell cytoplasm. When levels of
PER and CRY reach a threshold, they form heterodimers, feed back to the cell nucleus, and
negatively regulate CLOCK: BMAL1-mediated transcription of their own genes. Levels of PER are
regulated by casein kinase 1 epsilon (CK1ε) which phosphorylates these proteins and marks them
for degradation, thereby appropriately delaying negative feedback. Whereas Clock is constitutively
expressed, a secondary feedback loop drives the transcription of ROR and Rev-Erbα that, in turn,
induce rhythms in Bmal1 transcription through stimulatory and inhibitory actions on ROR response
elements (RRE) in the Bmal1 promotor, respectively. Clock-controlled genes are tissue-specific
genes that are produced rhythmically by the CLOCK:BMAL1 complex but are not part of the
clockwork mechanism (i.e., do not feed back onto the clockwork)
sex steroid (e.g., progesterone, estradiol, and testosterone) synthesis and secretion
and gametogenesis, respectively. Sex steroids and gonadotropins feed back to the
hypothalamus and anterior pituitary to regulate HPG axis activity. Upstream of the
GnRH system, two key neuropeptides were identified around the turn of the millen-
nium that have pronounced inhibitory and stimulatory actions on GnRH neurons,
namely RFamide-related peptide 3 and kisspeptin, respectively. The discovery and
significance of these neuropeptides are described below.
16.3.2 Kisspeptin
regulators of the reproductive axis by the circadian timing system. The requirement
for both high estradiol concentrations and a circadian timing signal to initiate the LH
surge ensures appropriate oocyte maturation and maximal sexual motivation coin-
cide with the time of ovulation.
As indicated previously, lesions of the SCN or the severing of connections
between the SCN and the preoptic area results in acyclicity in rodents (Brown-
Grant and Raisman 1977; Wiegand et al. 1980; Wiegand and Terasawa 1982).
Furthermore, genetic impairments of the molecular clockwork also disrupt estrous
cycles and the LH surge in mice (Khan and Kauffman 2012; Chu et al. 2013;
Dolatshad et al. 2006; Miller et al. 2004). Although the SCN can support behavioral
rhythms via a diffusible signal, the SCN-derived signal that generates the LH surge
is neural; SCN-lesioned female hamsters receiving fetal SCN transplants that do not
form neural connections with the host brain exhibit rhythmic behavior but not an
LH surge (Lehman et al. 1987; Silver et al. 1996; Meyer-Bernstein et al. 1999). As
described further below, both direct SCN-GnRH neuronal communication and
indirect connections from the SCN to estrogen receptor-α (ERα)-expressing cell
phenotypes that positively and negatively regulate the GnRH system integrate
circadian and estradiol signaling to initiate the preovulatory LH surge and ovulation
(Fig. 16.2).
Fig. 16.2 Circadian control of the preovulatory LH surge and ovulation. Model of interactions
among the circadian, RFRP-3, and kisspeptin systems in control of the preovulatory LH surge and
ovulation. At the time of the LH surge, the SCN coordinates estradiol positive feedback (through
kisspeptin cell activation) with removal of estradiol negative feedback (through suppression of
RFRP-3 cells in the DMH). GnRH, RFRP-3, and AVPV kisspeptin cells all exhibit rhythms in
clock protein or gene expression to further temporal precision in this circuit. See text for further
details
system that integrate circadian and estradiol signaling (Herbison and Theodosis 1992;
Hrabovszky et al. 2000; Wintermantel et al. 2006). As described below, RFRP-3 and
kisspeptin have emerged as key, estrogen-responsive cell phenotypes upstream of the
GnRH system that mediate circadian-controlled estrogen negative and positive
feedback, respectively.
Early findings pointed to the AVPV as a likely neural locus regulating estradiol
positive feedback and the LH surge. ERα expressing neurons the AVPV send
monosynaptic projections to GnRH neurons, express FOS at the time of the LH
surge, and AVPV lesions result in the loss of estrous cyclicity in intact and
ovariectomized/estradiol-treated rats (Wiegand and Terasawa 1982; Wiegand et al.
1980; Ronnekleiv and Kelly 1988; Gu and Simerly 1997; Le et al. 1999). However,
422 N. Gotlieb et al.
until relatively recently, the cell phenotype participating in estradiol positive feed-
back was unknown.
Following the discovery of kisspeptin, it was soon shown that ERα is expressed
in the majority of AVPV and Arc kisspeptin cells (Smith et al. 2005a; Dubois et al.
2015), with AVPV kisspeptin cells mediating estradiol positive feedback and the
Arc population negative feedback and GnRH pulsatility (Millar et al. 2010; Moore
et al. 2018). The SCN sends monosynaptic AVPergic projections to AVPV estrogen-
responsive kisspeptin cells (Fig. 16.2) that, in turn, project to kisspeptin-receptor-
expressing GnRH neurons in the POA to positively drive the LH surge (Vida et al.
2010; Chassard et al. 2015; Dubois et al. 2015; Piet et al. 2015; Simonneaux and
Bahougne 2015; Smarr et al. 2013; Smith et al. 2006; Williams et al. 2011; Yip et al.
2015). Significantly, the increased neural firing of kisspeptin cells in response to
AVP is estrogen dependent, consistent with a role of kisspeptin cells in integrating
estrogenic and circadian signaling (Piet et al. 2015).
In SCN-lesioned rats and Clock mutant mice, central administration of AVP
produces surge-like levels of LH (Palm et al. 1999; Miller et al. 2006). Importantly,
exogenous AVP only stimulates LH release when administered in the afternoon, the
timepoint when the preovulatory LH surge occurs in rodents (Palm et al. 2001).
Given this time-dependent sensitivity, we asked if time dependence of the surge was
due to daily changes in sensitivity of kisspeptin cells to AVP stimulation, daily
changes in GnRH cell sensitivity to kisspeptin stimulation, or a combination of both
processes. In Syrian hamsters, AVPV kisspeptin cells do not demonstrate time-
dependent changes in response to AVP stimulation, whereas GnRH neurons exhibit
daily changes in their responsiveness to kisspeptin, suggestive of autonomous
circadian timekeeping in cells downstream of the SCN (Williams et al. 2011).
Indeed, circadian oscillations in core clock genes that drive circadian rhythms at
the cellular level are found both in vitro and in vivo in GnRH cells (Chappell et al.
2003; Olcese et al. 2003; Zhao and Kriegsfeld 2009). Immortalized GnRH neurons
exhibit circadian rhythms in responsiveness to VIP and kisspeptin stimulation,
further indicating that the GnRH system maintains circadian timing potentially as
a mechanism mediating daily changes in responsiveness to upstream signaling (Zhao
and Kriegsfeld 2009). Despite not exhibiting daily changes to AVP stimulation,
kisspeptin cells express the clock gene Per1 and the AVPV exhibits sustained
circadian rhythms in clock gene expression in cultured AVPV explants (Jacobs
et al. 2016; Chassard et al. 2015). Whether these sustained rhythms in AVPV
kisspeptin cells confer daily changes in responsiveness to upstream neurochemicals
other than AVP remains to be determined.
A specific role for ARC kisspeptin cells in the generation of the preovulatory LH
surge has not been established. However, several findings suggest a role for Arc
kisspeptin cells in this process. For example, ablation of ARC kisspeptin cells results
in abnormal estrous cycles and LH surges, indicating a potential role in generating
the LH surge (Helena et al. 2015; Mittelman-Smith et al. 2016). Furthermore, this
population expresses receptors for both AVP and VIP (Lukas et al. 2010; Mounien
et al. 2006; Ronnekleiv et al. 2014), and AVP and VIP increase intracellular calcium
in subsets of Arc kisspeptin neurons in sexually dimorphic manner (Schafer et al.
2018).
16 Development and Modulation of Female Reproductive Function by Circadian. . . 423
As mentioned previously, prior to the LH surge and ovulation, estradiol acts via
negative feedback to maintain gonadotropins at low concentrations. Several lines of
evidence indicate a role for RFRP-3 cells in integrating circadian and estrogenic
signals to mediate estradiol negative feedback. First, RFRP-3 cells exhibit a high
activational state (as measured by FOS expression) during diestrus, reduced activity
around the time of the LH surge, and increased activity soon thereafter (Gibson et al.
2008). This pattern of timing is mediated by the SCN, with both AVP/VIPergic SCN
cell terminals forming close appositions with RFRP-3 cells in Syrian hamsters
(Gibson et al. 2008). Likewise, RFRP-3 neurons express ERα and increase their
activity in response to estradiol injections at the time of the LH surge (Kriegsfeld
et al. 2006; Molnar et al. 2011; Iwasa et al. 2012; Poling et al. 2012). Additionally,
analogous to effects seen for GnRH cells, VIP suppresses RFRP-3 cellular activity
around the time of the LH surge, but not prior to the surge, suggesting that removal
of RFPR-3-mediated estradiol negative feedback is accomplished via time-
dependent sensitivity to VIP signaling (Russo et al. 2015). This time-dependent
sensitivity is associated with rhythmic clock protein expression in RFRP-3 cells
(Russo et al. 2015). Additionally, RFRP-3 cell terminals form close appositions with
GnRH cell soma and terminals at median eminence, both of which express Gpr147
(Kriegsfeld et al. 2006; Johnson et al. 2007; Bentley et al. 2010; Kriegsfeld et al.
2010). The same impact of RFRP-3 is observed in both an estradiol surge implant
model (Gibson et al. 2008) and during the afternoon of proestrus (Henningsen et al.
2017). Finally, in mice, RFRP-3 similarly inhibits LH secretion when administered
at the time of the preovulatory LH surge, but not during diestrus (Ancel et al. 2017).
Together, these findings suggest that RFRP-3 cells integrate estrogen and circadian
signaling to time the removal of estradiol negative feedback with stimulation of the
LH surge.
daily surges during pregnancy (a diurnal and nocturnal surge) that maintains the
viability of the corpora lutea (CL) and the secretion of progesterone in the first half of
pregnancy. Around midpregnancy, these surges cease and placental lactogens main-
tain progesterone secretion for the remainder of gestation (Robertson and Friesen
1981; Bertram et al. 2010; Freeman et al. 1974; Shiu et al. 1973; Strauss et al. 1996).
The SCN regulates prolactin release through the pacing of inhibiting and stimulating
factors for this hormone, principally dopamine (DA) and oxytocin, respectively
(Bertram et al. 2010; Freeman et al. 2000; Mai et al. 1994). This regulation likely
occurs via VIPergic projections from the SCN, as both arcuate tuberoinfundibular
DA (TIDA) neurons and paraventricular nucleus (PVN) oxytocin neurons are
innervated by VIPergic fibers originating in the SCN (Arey and Freeman 1992;
Teclemariam-Mesbah et al. 1997; Gerhold et al. 2001; Egli et al. 2004; Poletini et al.
2010). In addition, VIP antisense oligonucleotides aimed at the SCN abolish prolac-
tin surges in rats (Harney et al. 1996). As would be expected given this mechanism
of control, SCN lesions abolish prolactin surges (Kawakami et al. 1980; Pan and
Gala 1985). Likewise, prolactin surges entrain to light:dark cycles and free run in
constant darkness (Bethea and Neill 1979; Yogev and Terkel 1980). Finally, knock-
down of essential clock genes (Per1, Per2, and Clock) in the SCN abolish prolactin
surges (Poletini et al. 2007), and mice lacking a functional Clock gene exhibit
reduced concentrations of progesterone and marked pregnancy failure, suggestive
of disrupted prolactin timing (Miller et al. 2004). These findings point to the critical
role of precisely timed prolactin secretion for the maintenance of pregnancy.
In addition to SCN regulation of TIDA and oxytocin neurons, kisspeptin cells
project to TIDA neurons and kisspeptin administration increases prolactin release
via dopaminergic cell inhibition (Szawka et al. 2010; Sawai et al. 2012, 2014; Saedi
et al. 2018). Furthermore, kisspeptin neurons express prolactin receptors, and
hyperprolactinemia causes reduced activity of hypothalamic kisspeptin cells,
indicating reciprocal communication between kisspeptin and TIDA neurons (Saedi
et al. 2018; Sonigo et al. 2012). As with TIDA and oxytocin regulation by the SCN,
arcuate kisspeptin cells express VIP receptor and VIP administration alters their
cellular activity (Schafer et al. 2018), suggesting that the SCN likely regulates TIDA
neurons directly and indirectly, via kisspeptin cells through VIPergic signaling.
Whether or not disruptions to kisspeptin cell timing affect prolactin release and
pregnancy success remains to be determined.
In contrast to rodents, the pituitary gland (or pituitary hormones) is not required
for the initiation and maintenance of pregnancy in humans. In the first 8 weeks of
pregnancy, the CL is maintained by human chorionic gonadotropin (hCG) produced
by the trophoblast (a layer of tissue that later forms part of the placenta), with
placental progesterone sufficient to maintain pregnancy thereafter (Malassine et al.
2003). Although humans do not exhibit a twice daily prolactin surge and, progester-
one secretion is maintained via the placenta, circadian disruptions profoundly reduce
pregnancy success as seen in rodents (Gamble et al. 2013; Bonzini et al. 2011;
Knutsson 2003; Lin et al. 2011). Given that the placenta is responsible for fetal-
maternal nutrient exchange and maintains rhythmic clock gene and hormonal
expression, it is likely that circadian disruption negatively influences pregnancy
16 Development and Modulation of Female Reproductive Function by Circadian. . . 425
success in humans via disruptions to this structure, a hypothesis that has not yet been
explored.
Circadian disruptions that alter maternal endocrine timing signals can impair
maternal–fetal synchronization and fetal development (Fig. 16.3). A major maternal
signal providing photoperiodic information to the embryo is melatonin, a hor-
mone secreted from the maternal pineal gland at darkness. Melatonin crosses the
placenta and influences the embryo’s physiological rhythms and development
(Schenker et al. 1998; Williams et al. 1991; Okatani et al. 1998). Melatonin receptors
are widespread in the fetal nervous system (including the SCN), as well as in
peripheral organs, beginning early in fetal development (Drew et al. 1998; Thomas
et al. 2002; Rivkees and Reppert 1991; Vanecek 1988; Torres-Farfan et al. 2008).
Likewise, the placenta expresses melatonin receptors, and melatonin plays a role in
placenta development (Tamura et al. 2008). Nighttime melatonin secretion increases
throughout pregnancy (Nakamura et al. 2001), and melatonin rhythms in maternal
blood are mirrored in fetal circulation (Kennaway et al. 1996; Okatani et al. 1998).
In addition to melatonin, fetal rhythms are also entrained by maternal cortisol and
body temperature rhythms (Seron-Ferre et al. 2001), and possibly feeding times
(Novakova et al. 2010) (Fig. 16.3).
Because the pineal gland matures only after birth and the developing fetus and
newborn do not produce their own melatonin, maternal melatonin signaling is
conveyed to offspring through the placenta (in utero) and milk (after birth) and is
required for fetal/newborn rhythms. Thus, it is not surprising that maternal circadian
disruptions are reflected in fetal rhythms. In nonhuman primates, for example,
maternal light exposure disrupts rhythmic expression of fetal clock genes that can
be rescued by maternal melatonin administration (Torres-Farfan et al. 2006). Like-
wise, suppression of maternal melatonin by constant light exposure during preg-
nancy is associated with intrauterine growth restriction (IUGR), lower
concentrations and altered rhythms of cortisol, modified mRNA expression of
16 Development and Modulation of Female Reproductive Function by Circadian. . . 427
Fig. 16.3 Maternal–fetal rhythm synchronization. The developing fetus is exposed to an array of
time cues from its mother. The main signal providing rhythmic information is melatonin, secreted
from the mother’s pineal gland at darkness and crossing the placenta to the embryo. Additional
entraining signals include maternal cortisol, body temperature rhythms, and feeding times. In
addition, the developing embryo is exposed to a rhythmic environment via clock genes that are
rhythmically expressed in maternal reproductive tissues, including the oviduct, uterus, and placenta.
This rhythmic environment is likely critical for normal fetal and postnatal development. Circadian
disruptions that alter maternal endocrine timing signals can impair maternal–fetal synchronization
and fetal development. Clocks indicate rhythmic expression of clock genes
clock genes and clock-controlled genes in the fetal adrenal gland, and aberrant
adrenal response to adrenocorticotropic hormone (ACTH) in rats. Melatonin admin-
istration during the subjective night rescues all of the above (Mendez et al. 2012). In
in vitro fertilization (IVF) treatments, melatonin promotes embryo development
(Ishizuka et al. 2000; Tian et al. 2010; Sampaio et al. 2012). Likewise, melatonin
administration prior to IVF treatments and throughout pregnancy is associated with
improved pregnancy outcomes; fertility rates are 50% higher in melatonin treated
IVF cycles (Rizzo et al. 2010; Unfer et al. 2011). These findings raise questions
regarding the static nature of the environment that the conceptus is typically exposed
428 N. Gotlieb et al.
to in cultured IVF conditions and whether a rhythmic environment that better mimics
the in vivo milieu would increase IVF success rates. Likewise, the timing of IVF
embryo implantation may be important to treatment success, but this remains to be
determined. Nonetheless, the dependence of fetal/early postnatal rhythms on mater-
nal environment further underscores the importance of maintaining circadian health
during pregnancy and lactation.
The circadian timing system participates in all aspects of female reproduction, from
ovulation to childbirth. Disruptions to circadian timing have marked negative
consequences for ovulation, pregnancy success and maintenance, and offspring
development. The modern world makes circadian disruption virtually inescapable,
underscoring the importance of developing safe and effective strategies to maximize
circadian health in the face of pervasive circadian disruption. Likewise, educating
couples about the importance of circadian health represents an important consider-
ation for health professionals providing counseling during family planning.
Key References
Endo and Watanabe (1989)—This article describes the effects of non-24 h days on
female reproduction in mice. The authors found that mice maintained on a 22 or
26 h light:dark cycle exhibit decreased mating behavior and experience higher
rates of fetal resorption, reduced embryo weights, and delayed development.
Kriegsfeld et al. (2006)—Based on the discovery of gonadotropin-inhibitory hor-
mone (GnIH) in birds, the authors provided the first, broad characterization of
GnIH (also known as RFamide-related peptide-3) in mammals (i.e., rats, mice
and hamsters), establishing a more general modulatory role for this neuropeptide
across species.
Mendez et al. (2012)—This article demonstrated a role for melatonin in fetal
development and rhythm generation/synchronization. The authors established
that exposure of pregnant rats to constant light during the second half of gestation
negatively affects fetal development, with embryos exhibiting intrauterine growth
retardation and disrupted adrenal clock genes and clock-controlled genes as well
as altered corticosterone rhythms. However, when mothers received melatonin
during their subjective night, reproductive function was rescued.
Miller et al. (2006)—This article examined reproduction of Clock mutant mice,
establishing an important role for circadian rhythms in female reproductive
function. The authors found that Clock mutant mice exhibit disruptions in estrous
cyclicity and pregnancy maintenance, including higher rates of fetal resorptions
and fewer pregnancies reaching full term.
Seminara et al. (2003)—This article employed complementary approaches in
humans and mice to investigate the onset of puberty. The authors found that
mutations in the G-protein coupled receptor, GPR54, the cognate receptor
kisspeptin, caused idiopathic hypophysiotropic hypogonadism in both mice and
humans. This finding was key to identifying a critical role for kisspeptin in
reproductive functioning.
Takayama et al. (2003)—This article demonstrated a role for melatonin in timing
parturition. The authors found that female rats lacking melatonin deliver pups
throughout the day and night, unlike melatonin-proficient rats who enter parturi-
tion during the day. Melatonin administration to rats exclusively during the dark
phase restored parturition to daytime.
Tsutsui et al. (2000)—This article identified a hypothalamic peptide that inhibits
gonadotropin release in a vertebrate pituitary and named it gonadotropin-
inhibitory hormone (GnIH). This finding represented a discovery with broad
implications for understanding reproductive axis regulation.
References
Abe M, Herzog ED, Yamazaki S, Straume M, Tei H, Sakaki Y, Menaker M, Block GD (2002)
Circadian rhythms in isolated brain regions. J Neurosci 22(1):350–356
Adachi S, Yamada S, Takatsu Y, Matsui H, Kinoshita M, Takase K, Sugiura H, Ohtaki T,
Matsumoto H, Uenoyama Y, Tsukamura H, Inoue K, Maeda K (2007) Involvement of
anteroventral periventricular metastin/kisspeptin neurons in estrogen positive feedback action
on luteinizing hormone release in female rats. J Reprod Dev 53(2):367–378
432 N. Gotlieb et al.
Ahlborg G Jr, Axelsson G, Bodin L (1996) Shift work, nitrous oxide exposure and subfertility
among Swedish midwives. Int J Epidemiol 25(4):783–790
Akiyama S, Ohta H, Watanabe S, Moriya T, Hariu A, Nakahata N, Chisaka H, Matsuda T,
Kimura Y, Tsuchiya S, Tei H, Okamura K, Yaegashi N (2010) The uterus sustains stable
biological clock during pregnancy. Tohoku J Exp Med 221(4):287–298
Alvarez JD, Hansen A, Ord T, Bebas P, Chappell PE, Giebultowicz JM, Williams C, Moss S,
Sehgal A (2008) The circadian clock protein BMAL1 is necessary for fertility and proper
testosterone production in mice. J Biol Rhythm 23(1):26–36. https://doi.org/10.1177/
0748730407311254
Ancel C, Bentsen AH, Sebert ME, Tena-Sempere M, Mikkelsen JD, Simonneaux V (2012) Stimu-
latory effect of RFRP-3 on the gonadotrophic axis in the male Syrian hamster: the exception
proves the rule. Endocrinology 153(3):1352–1363. https://doi.org/10.1210/en.2011-1622
Ancel C, Inglis MA, Anderson GM (2017) Central RFRP-3 stimulates LH secretion in male mice
and has cycle stage-dependent inhibitory effects in females. Endocrinology 158(9):2873–2883.
https://doi.org/10.1210/en.2016-1902
Arey BJ, Freeman ME (1992) Activity of vasoactive intestinal peptide and serotonin in the
paraventricular nucleus reflects the periodicity of the endogenous stimulatory rhythm regulating
prolactin secretion. Endocrinology 131(2):736–742. https://doi.org/10.1210/endo.131.2.
1639019
Bedrosian TA, Nelson RJ (2013) Influence of the modern light environment on mood. Mol
Psychiatry 18(7):751–757. https://doi.org/10.1038/mp.2013.70mp201370
Bentley GE, Tsutsui K, Kriegsfeld LJ (2010) Recent studies of gonadotropin-inhibitory hormone
(GnIH) in the mammalian hypothalamus, pituitary and gonads. Brain Res 1364:62–71. https://
doi.org/10.1016/j.brainres.2010.10.001
Berson DM, Dunn FA, Takao M (2002) Phototransduction by retinal ganglion cells that set the
circadian clock. Science 295(5557):1070–1073. https://doi.org/10.1126/science.1067262
Bertram R, Helena CV, Gonzalez-Iglesias AE, Tabak J, Freeman ME (2010) A tale of two rhythms:
the emerging roles of oxytocin in rhythmic prolactin release. J Neuroendocrinol 22(7):778–784.
https://doi.org/10.1111/j.1365-2826.2010.02012.x
Bethea CL, Neill JD (1979) Prolactin secretion after cervical stimulation of rats maintained in
constant dark or constant light. Endocrinology 104(4):870–876. https://doi.org/10.1210/endo-
104-4-870
Blaustein JD, Tetel MJ, Ricciardi KH, Delville Y, Turcotte JC (1994) Hypothalamic ovarian steroid
hormone-sensitive neurons involved in female sexual behavior. Psychoneuroendocrinology 19
(5–7):505–516
Boden MJ, Varcoe TJ, Voultsios A, Kennaway DJ (2010) Reproductive biology of female Bmal1
null mice. Reproduction 139(6):1077–1090. https://doi.org/10.1530/REP-09-0523
Boden MJ, Varcoe TJ, Kennaway DJ (2013) Circadian regulation of reproduction: from gamete to
offspring. Prog Biophys Mol Biol 113(3):387–397. https://doi.org/10.1016/j.pbiomolbio.2013.
01.003
Boer K, Lincoln DW, Swaab DF (1975) Effects of electrical stimulation of the neurohypophysis on
labour in the rat. J Endocrinol 65(2):163–176
Bonini JA, Jones KA, Adham N, Forray C, Artymyshyn R, Durkin MM, Smith KE, Tamm JA,
Boteju LW, Lakhlani PP, Raddatz R, Yao WJ, Ogozalek KL, Boyle N, Kouranova EV, Quan Y,
Vaysse PJ, Wetzel JM, Branchek TA, Gerald C, Borowsky B (2000) Identification and
characterization of two G protein-coupled receptors for neuropeptide FF. J Biol Chem 275
(50):39324–39331. https://doi.org/10.1074/jbc.M004385200
Bonzini M, Palmer KT, Coggon D, Carugno M, Cromi A, Ferrario MM (2011) Shift work and
pregnancy outcomes: a systematic review with meta-analysis of currently available epidemio-
logical studies. BJOG 118(12):1429–1437. https://doi.org/10.1111/j.1471-0528.2011.03066.x
Bosc MJ, Nicolle A (1980) Influence of photoperiod on the time of parturition in the rat. I. Effect
of the length of daily illumination on normal or adrenalectomized animals. Reprod Nutr Dev
20(3A):735–745
16 Development and Modulation of Female Reproductive Function by Circadian. . . 433
Deng N, Kohn TP, Lipshultz LI, Pastuszak AW (2018) The relationship between shift work and
Men’s health. Sex Med Rev 6(3):446–456. https://doi.org/10.1016/j.sxmr.2017.11.009
Dolatshad H, Campbell EA, O’Hara L, Maywood ES, Hastings MH, Johnson MH (2006) Devel-
opmental and reproductive performance in circadian mutant mice. Hum Reprod 21(1):68–79.
https://doi.org/10.1093/humrep/dei313
Dominguez Rubio AP, Sordelli MS, Salazar AI, Aisemberg J, Bariani MV, Cella M, Rosenstein
RE, Franchi AM (2014) Melatonin prevents experimental preterm labor and increases offspring
survival. J Pineal Res 56(2):154–162. https://doi.org/10.1111/jpi.12108
Drew JE, Williams LM, Hannah LT, Barrett P, Abramovich DR (1998) Melatonin receptors in the
human fetal kidney: 2-[125I]iodomelatonin binding sites correlated with expression of Mel1a
and Mel1b receptor genes. J Endocrinol 156(2):261–267
Dubois SL, Acosta-Martinez M, DeJoseph MR, Wolfe A, Radovick S, Boehm U, Urban JH, Levine
JE (2015) Positive, but not negative feedback actions of estradiol in adult female mice require
estrogen receptor alpha in kisspeptin neurons. Endocrinology 156(3):1111–1120. https://doi.
org/10.1210/en.2014-1851
Ducret E, Anderson GM, Herbison AE (2009) RFamide-related peptide-3, a mammalian
gonadotropin-inhibitory hormone ortholog, regulates gonadotropin-releasing hormone neuron
firing in the mouse. Endocrinology 150(6):2799–2804. https://doi.org/10.1210/en.2008-1623
Edwards RG, Steptoe PC, Fowler RE, Baillie J (1980) Observations on preovulatory human ovarian
follicles and their aspirates. Br J Obstet Gynaecol 87(9):769–779
Egli M, Bertram R, Sellix MT, Freeman ME (2004) Rhythmic secretion of prolactin in rats: action
of oxytocin coordinated by vasoactive intestinal polypeptide of suprachiasmatic nucleus origin.
Endocrinology 145(7):3386–3394. https://doi.org/10.1210/en.2003-1710
Elkind-Hirsch K, Ravnikar V, Tulchinsky D, Schiff I, Ryan KJ (1984) Episodic secretory patterns
of immunoreactive luteinizing hormone-releasing hormone (IR-LH-RH) in the systemic circu-
lation of normal women throughout the menstrual cycle. Fertil Steril 41(1):56–61
Endo A, Watanabe T (1989) Effects of non-24-hour days on reproductive efficacy and embryonic
development in mice. Gamete Res 22(4):435–441. https://doi.org/10.1002/mrd.1120220409
Fernandez RC, Marino JL, Varcoe TJ, Davis S, Moran LJ, Rumbold AR, Brown HM, Whitrow MJ,
Davies MJ, Moore VM (2016) Fixed or rotating night shift work undertaken by women:
implications for fertility and miscarriage. Semin Reprod Med 34(2):74–82. https://doi.org/10.
1055/s-0036-1571354
Fonken LK, Nelson RJ (2014) The effects of light at night on circadian clocks and metabolism.
Endocr Rev 35, 648–670. https://doi.org/10.1210/er.2013-1051
Foradori CD, Coolen LM, Fitzgerald ME, Skinner DC, Goodman RL, Lehman MN (2002)
Colocalization of progesterone receptors in parvicellular dynorphin neurons of the ovine
preoptic area and hypothalamus. Endocrinology 143(11):4366–4374. https://doi.org/10.1210/
en.2002-220586
Francl JM, Kaur G, Glass JD (2010) Regulation of vasoactive intestinal polypeptide release in the
suprachiasmatic nucleus circadian clock. Neuroreport 21(16):1055–1059. https://doi.org/10.
1097/WNR.0b013e32833fcba4
Freeman ME, Smith MS, Nazian SJ, Neill JD (1974) Ovarian and hypothalamic control of the daily
surges of prolactin secretion during pseudopregnancy in the rat. Endocrinology 94(3):875–882.
https://doi.org/10.1210/endo-94-3-875
Freeman ME, Kanyicska B, Lerant A, Nagy G (2000) Prolactin: structure, function, and regulation
of secretion. Physiol Rev 80(4):1523–1631. https://doi.org/10.1152/physrev.2000.80.4.1523
Funabashi T, Aiba S, Sano A, Shinohara K, Kimura F (1999) Intracerebroventricular injection of
arginine-vasopressin V1 receptor antagonist attenuates the surge of luteinizing hormone and
prolactin secretion in proestrous rats. Neurosci Lett 260(1):37–40
Gamble KL, Resuehr D, Johnson CH (2013) Shift work and circadian dysregulation of reproduc-
tion. Front Endocrinol 4:92. https://doi.org/10.3389/fendo.2013.00092
George JT, Hendrikse M, Veldhuis JD, Clarke IJ, Anderson RA, Millar RP (2017) Effect of
gonadotropin-inhibitory hormone on luteinizing hormone secretion in humans. Clin Endocrinol
86(5):731–738. https://doi.org/10.1111/cen.13308
16 Development and Modulation of Female Reproductive Function by Circadian. . . 435
Gerhold LM, Horvath TL, Freeman ME (2001) Vasoactive intestinal peptide fibers innervate
neuroendocrine dopaminergic neurons. Brain Res 919(1):48–56
Germain AM, Valenzuela GJ, Ivankovic M, Ducsay CA, Gabella C, Seron-Ferre M (1993)
Relationship of circadian rhythms of uterine activity with term and preterm delivery. Am J
Obstet Gynecol 168(4):1271–1277
Gibson EM, Humber SA, Jain S, Williams WP 3rd, Zhao S, Bentley GE, Tsutsui K, Kriegsfeld LJ
(2008) Alterations in RFamide-related peptide expression are coordinated with the preovulatory
luteinizing hormone surge. Endocrinology 149(10):4958–4969. https://doi.org/10.1210/en.
2008-0316
Gibson EM, Wang C, Tjho S, Khattar N, Kriegsfeld LJ (2010) Experimental ‘jet lag’ inhibits adult
neurogenesis and produces long-term cognitive deficits in female hamsters. PLoS One 5(12):
e15267. https://doi.org/10.1371/journal.pone.0015267
Goodman RL, Lehman MN, Smith JT, Coolen LM, de Oliveira CV, Jafarzadehshirazi MR,
Pereira A, Iqbal J, Caraty A, Ciofi P, Clarke IJ (2007) Kisspeptin neurons in the arcuate nucleus
of the ewe express both dynorphin A and neurokinin B. Endocrinology 148(12):5752–5760.
https://doi.org/10.1210/en.2007-0961
Goodman RL, Hileman SM, Nestor CC, Porter KL, Connors JM, Hardy SL, Millar RP, Cernea M,
Coolen LM, Lehman MN (2013) Kisspeptin, neurokinin B, and dynorphin act in the arcuate
nucleus to control activity of the GnRH pulse generator in ewes. Endocrinology 154
(11):4259–4269. https://doi.org/10.1210/en.2013-1331
Goodman RL, Coolen LM, Lehman MN (2014) A role for neurokinin B in pulsatile GnRH
secretion in the ewe. Neuroendocrinology 99(1):18–32. https://doi.org/10.1159/000355285
Gouarderes C, Mazarguil H, Mollereau C, Chartrel N, Leprince J, Vaudry H, Zajac JM (2007)
Functional differences between NPFF1 and NPFF2 receptor coupling: high intrinsic activities of
RFamide-related peptides on stimulation of [35S]GTPgammaS binding. Neuropharmacology
52(2):376–386. https://doi.org/10.1016/j.neuropharm.2006.07.034
Gu GB, Simerly RB (1997) Projections of the sexually dimorphic anteroventral periventricular
nucleus in the female rat. J Comp Neurol 384(1):142–164
Guran T, Tolhurst G, Bereket A, Rocha N, Porter K, Turan S, Gribble FM, Kotan LD, Akcay T,
Atay Z, Canan H, Serin A, O’Rahilly S, Reimann F, Semple RK, Topaloglu AK (2009)
Hypogonadotropic hypogonadism due to a novel missense mutation in the first extracellular
loop of the neurokinin B receptor. J Clin Endocrinol Metab 94(10):3633–3639. https://doi.org/
10.1210/jc.2009-0551
Harney JP, Scarbrough K, Rosewell KL, Wise PM (1996) In vivo antisense antagonism of
vasoactive intestinal peptide in the suprachiasmatic nuclei causes aging-like changes in the
estradiol-induced luteinizing hormone and prolactin surges. Endocrinology 137(9):3696–3701.
https://doi.org/10.1210/endo.137.9.8756535
Hastings M, O’Neill JS, Maywood ES (2007) Circadian clocks: regulators of endocrine and
metabolic rhythms. J Endocrinol 195(2):187–198. https://doi.org/10.1677/JOE-07-0378
Hattar S, Liao HW, Takao M, Berson DM, Yau KW (2002) Melanopsin-containing retinal ganglion
cells: architecture, projections, and intrinsic photosensitivity. Science 295(5557):1065–1070.
https://doi.org/10.1126/science.1069609
Helena CV, Toporikova N, Kalil B, Stathopoulos AM, Pogrebna VV, Carolino RO, Anselmo-
Franci JA, Bertram R (2015) KNDy neurons modulate the magnitude of the steroid-induced
luteinizing hormone surges in ovariectomized rats. Endocrinology 156(11):4200–4213. https://
doi.org/10.1210/en.2015-1070
Henningsen JB, Ancel C, Mikkelsen JD, Gauer F, Simonneaux V (2017) Roles of RFRP-3 in the
daily and seasonal regulation of reproductive activity in female Syrian hamsters. Endocrinology
158(3):652–663. https://doi.org/10.1210/en.2016-1689
Herbison AE, Theodosis DT (1992) Localization of oestrogen receptors in preoptic neurons
containing neurotensin but not tyrosine hydroxylase, cholecystokinin or luteinizing hormone-
releasing hormone in the male and female rat. Neuroscience 50(2):283–298
436 N. Gotlieb et al.
Kalsbeek A, Buijs RM, Engelmann M, Wotjak CT, Landgraf R (1995) In vivo measurement of a
diurnal variation in vasopressin release in the rat suprachiasmatic nucleus. Brain Res 682
(1–2):75–82
Karsch FJ, Bowen JM, Caraty A, Evans NP, Moenter SM (1997) Gonadotropin-releasing hormone
requirements for ovulation. Biol Reprod 56(2):303–309
Kawakami M, Arita J, Yoshioka E (1980) Loss of estrogen-induced daily surges of prolactin and
gonadotropins by suprachiasmatic nucleus lesions in ovariectomized rats. Endocrinology 106
(4):1087–1092. https://doi.org/10.1210/endo-106-4-1087
Kennaway DJ, Goble FC, Stamp GE (1996) Factors influencing the development of melatonin
rhythmicity in humans. J Clin Endocrinol Metab 81(4):1525–1532. https://doi.org/10.1210/
jcem.81.4.8636362
Kennaway DJ, Varcoe TJ, Mau VJ (2003) Rhythmic expression of clock and clock-controlled
genes in the rat oviduct. Mol Hum Reprod 9(9):503–507
Kerdelhue B, Brown S, Lenoir V, Queenan JT Jr, Jones GS, Scholler R, Jones HW Jr (2002) Timing
of initiation of the preovulatory luteinizing hormone surge and its relationship with the circadian
cortisol rhythm in the human. Neuroendocrinology 75(3):158–163
Khan AR, Kauffman AS (2012) The role of kisspeptin and RFamide-related peptide-3 neurones in
the circadian-timed preovulatory luteinising hormone surge. J Neuroendocrinol 24(1):131–143.
https://doi.org/10.1111/j.1365-2826.2011.02162.x
Kivela A, Kauppila A, Leppaluoto J, Vakkuri O (1989) Serum and amniotic fluid melatonin during
human labor. J Clin Endocrinol Metab 69(5):1065–1068. https://doi.org/10.1210/jcem-69-5-1065
Knutsson A (2003) Health disorders of shift workers. Occup Med (Lond) 53(2):103–108
Kotani M, Detheux M, Vandenbogaerde A, Communi D, Vanderwinden JM, Le Poul E,
Brezillon S, Tyldesley R, Suarez-Huerta N, Vandeput F, Blanpain C, Schiffmann SN,
Vassart G, Parmentier M (2001) The metastasis suppressor gene KiSS-1 encodes kisspeptins,
the natural ligands of the orphan G protein-coupled receptor GPR54. J Biol Chem 276
(37):34631–34636. https://doi.org/10.1074/jbc.M104847200
Kouba AJ, Burkhardt BR, Alvarez IM, Goodenow MM, Buhi WC (2000) Oviductal plasminogen
activator inhibitor-1 (PAI-1): mRNA, protein, and hormonal regulation during the estrous cycle
and early pregnancy in the pig. Mol Reprod Dev 56(3):378–386. https://doi.org/10.1002/1098-
2795(200007)56:3<378::AID-MRD8>3.0.CO;2-B
Kovanen L, Saarikoski ST, Aromaa A, Lonnqvist J, Partonen T (2010) ARNTL (BMAL1) and
NPAS2 gene variants contribute to fertility and seasonality. PLoS One 5(4):e10007. https://doi.
org/10.1371/journal.pone.0010007
Krajewski SJ, Burke MC, Anderson MJ, McMullen NT, Rance NE (2010) Forebrain projections of
arcuate neurokinin B neurons demonstrated by anterograde tract-tracing and monosodium
glutamate lesions in the rat. Neuroscience 166(2):680–697. https://doi.org/10.1016/j.neurosci
ence.2009.12.053
Kriegsfeld LJ, Silver R, Gore AC, Crews D (2002) Vasoactive intestinal polypeptide contacts on
gonadotropin-releasing hormone neurones increase following puberty in female rats. J
Neuroendocrinol 14(9):685–690
Kriegsfeld LJ, Mei DF, Bentley GE, Ubuka T, Mason AO, Inoue K, Ukena K, Tsutsui K, Silver R
(2006) Identification and characterization of a gonadotropin-inhibitory system in the brains of
mammals. Proc Natl Acad Sci USA 103(7):2410–2415. https://doi.org/10.1073/pnas.
0511003103
Kriegsfeld LJ, Gibson EM, Williams WP 3rd, Zhao S, Mason AO, Bentley GE, Tsutsui K (2010)
The roles of RFamide-related peptide-3 in mammalian reproductive function and behaviour. J
Neuroendocrinol 22(7):692–700. https://doi.org/10.1111/j.1365-2826.2010.02031.x
Kriegsfeld LJ, Jennings KJ, Bentley GE, Tsutsui K (2018) Gonadotrophin-inhibitory hormone and
its mammalian orthologue RFamide-related peptide-3: discovery and functional implications for
reproduction and stress. J Neuroendocrinol 30(7):e12597. https://doi.org/10.1111/jne.12597
Lawson CC, Whelan EA, Lividoti Hibert EN, Spiegelman D, Schernhammer ES, Rich-Edwards JW
(2011) Rotating shift work and menstrual cycle characteristics. Epidemiology 22(3):305–312.
https://doi.org/10.1097/EDE.0b013e3182130016
438 N. Gotlieb et al.
Lawson CC, Rocheleau CM, Whelan EA, Lividoti Hibert EN, Grajewski B, Spiegelman D, Rich-
Edwards JW (2012) Occupational exposures among nurses and risk of spontaneous abortion.
Am J Obstet Gynecol 206(4):327.e321–328. https://doi.org/10.1016/j.ajog.2011.12.030
Le WW, Berghorn KA, Rassnick S, Hoffman GE (1999) Periventricular preoptic area neurons
coactivated with luteinizing hormone (LH)-releasing hormone (LHRH) neurons at the time of
the LH surge are LHRH afferents. Endocrinology 140(1):510–519. https://doi.org/10.1210/
endo.140.1.6403
Lee JH, Welch DR (1997) Suppression of metastasis in human breast carcinoma MDA-MB-435
cells after transfection with the metastasis suppressor gene, KiSS-1. Cancer Res 57
(12):2384–2387
Lee JH, Miele ME, Hicks DJ, Phillips KK, Trent JM, Weissman BE, Welch DR (1996) KiSS-1, a
novel human malignant melanoma metastasis-suppressor gene. J Natl Cancer Inst 88
(23):1731–1737
Lehman MN, Silver R, Gladstone WR, Kahn RM, Gibson M, Bittman EL (1987) Circadian
rhythmicity restored by neural transplant. Immunocytochemical characterization of the graft
and its integration with the host brain. J Neurosci 7(6):1626–1638
Levine JE (1997) New concepts of the neuroendocrine regulation of gonadotropin surges in rats.
Biol Reprod 56(2):293–302
Levine JE, Ramirez VD (1982) Luteinizing hormone-releasing hormone release during the rat
estrous cycle and after ovariectomy, as estimated with push-pull cannulae. Endocrinology 111
(5):1439–1448
Lin YC, Chen MH, Hsieh CJ, Chen PC (2011) Effect of rotating shift work on childbearing and
birth weight: a study of women working in a semiconductor manufacturing factory. World J
Pediatr 7(2):129–135. https://doi.org/10.1007/s12519-011-0265-9
Lucas RJ, Hattar S, Takao M, Berson DM, Foster RG, Yau KW (2003) Diminished pupillary light
reflex at high irradiances in melanopsin-knockout mice. Science 299(5604):245–247. https://
doi.org/10.1126/science.1077293
Lukas M, Bredewold R, Neumann ID, Veenema AH (2010) Maternal separation interferes with
developmental changes in brain vasopressin and oxytocin receptor binding in male rats.
Neuropharmacology 58(1):78–87. https://doi.org/10.1016/j.neuropharm.2009.06.020
Magiakou MA, Mastorakos G, Rabin D, Margioris AN, Dubbert B, Calogero AE, Tsigos C,
Munson PJ, Chrousos GP (1996) The maternal hypothalamic-pituitary-adrenal axis in the
third trimester of human pregnancy. Clin Endocrinol 44(4):419–428
Mahoney MM (2010) Shift work, jet lag, and female reproduction. Int J Endocrinol 2010:813764.
https://doi.org/10.1155/2010/813764
Mahoney MM, Sisk C, Ross HE, Smale L (2004) Circadian regulation of gonadotropin-releasing
hormone neurons and the preovulatory surge in luteinizing hormone in the diurnal rodent,
Arvicanthis niloticus, and in a nocturnal rodent, Rattus norvegicus. Biol Reprod 70
(4):1049–1054. https://doi.org/10.1095/biolreprod.103.021360
Mai LM, Shieh KR, Pan JT (1994) Circadian changes of serum prolactin levels and
tuberoinfundibular dopaminergic neuron activities in ovariectomized rats treated with or with-
out estrogen: the role of the suprachiasmatic nuclei. Neuroendocrinology 60(5):520–526.
https://doi.org/10.1159/000126789
Malassine A, Frendo JL, Evain-Brion D (2003) A comparison of placental development and
endocrine functions between the human and mouse model. Hum Reprod Update 9(6):531–539
Matsui H, Takatsu Y, Kumano S, Matsumoto H, Ohtaki T (2004) Peripheral administration of
metastin induces marked gonadotropin release and ovulation in the rat. Biochem Biophys Res
Commun 320(2):383–388. https://doi.org/10.1016/j.bbrc.2004.05.185
McEwen BS, Jones KJ, Pfaff DW (1987) Hormonal control of sexual behavior in the female rat:
molecular, cellular and neurochemical studies. Biol Reprod 36(1):37–45
Mendez N, Abarzua-Catalan L, Vilches N, Galdames HA, Spichiger C, Richter HG, Valenzuela GJ,
Seron-Ferre M, Torres-Farfan C (2012) Timed maternal melatonin treatment reverses circadian
disruption of the fetal adrenal clock imposed by exposure to constant light. PLoS One 7(8):
e42713. https://doi.org/10.1371/journal.pone.0042713
16 Development and Modulation of Female Reproductive Function by Circadian. . . 439
Menon R, Bonney EA, Condon J, Mesiano S, Taylor RN (2016) Novel concepts on pregnancy
clocks and alarms: redundancy and synergy in human parturition. Hum Reprod Update 22
(5):535–560. https://doi.org/10.1093/humupd/dmw022
Meyer-Bernstein EL, Jetton AE, Matsumoto SI, Markuns JF, Lehman MN, Bittman EL (1999)
Effects of suprachiasmatic transplants on circadian rhythms of neuroendocrine function in
golden hamsters. Endocrinology 140(1):207–218. https://doi.org/10.1210/endo.140.1.6428
Millar RP, Roseweir AK, Tello JA, Anderson RA, George JT, Morgan K, Pawson AJ (2010)
Kisspeptin antagonists: unraveling the role of kisspeptin in reproductive physiology. Brain Res
1364:81–89. https://doi.org/10.1016/j.brainres.2010.09.044
Miller BH, Takahashi JS (2013) Central circadian control of female reproductive function. Front
Endocrinol 4:195. https://doi.org/10.3389/fendo.2013.00195
Miller BH, Olson SL, Turek FW, Levine JE, Horton TH, Takahashi JS (2004) Circadian clock
mutation disrupts estrous cyclicity and maintenance of pregnancy. Curr Biol 14(15):1367–1373.
https://doi.org/10.1016/j.cub.2004.07.055
Miller BH, Olson SL, Levine JE, Turek FW, Horton TH, Takahashi JS (2006) Vasopressin
regulation of the proestrous luteinizing hormone surge in wild-type and clock mutant mice.
Biol Reprod 75(5):778–784. https://doi.org/10.1095/biolreprod.106.052845
Mittelman-Smith MA, Williams H, Krajewski-Hall SJ, Lai J, Ciofi P, McMullen NT, Rance NE
(2012) Arcuate kisspeptin/neurokinin B/dynorphin (KNDy) neurons mediate the estrogen
suppression of gonadotropin secretion and body weight. Endocrinology 153(6):2800–2812.
https://doi.org/10.1210/en.2012-1045
Mittelman-Smith MA, Krajewski-Hall SJ, McMullen NT, Rance NE (2016) Ablation of KNDy
neurons results in hypogonadotropic hypogonadism and amplifies the steroid-induced LH surge
in female rats. Endocrinology 157(5):2015–2027. https://doi.org/10.1210/en.2015-1740
Moenter SM, Brand RC, Karsch FJ (1992) Dynamics of gonadotropin-releasing hormone (GnRH)
secretion during the GnRH surge: insights into the mechanism of GnRH surge induction.
Endocrinology 130(5):2978–2984
Moline ML, Albers HE, Todd RB, Moore-Ede MC (1981) Light-dark entrainment of proestrous LH
surges and circadian locomotor activity in female hamsters. Horm Behav 15(4):451–458
Molnar CS, Kallo I, Liposits Z, Hrabovszky E (2011) Estradiol down-regulates RF-amide-related
peptide (RFRP) expression in the mouse hypothalamus. Endocrinology 152(4):1684–1690.
https://doi.org/10.1210/en.2010-1418
Mong JA, Pfaff DW (2003) Hormonal and genetic influences underlying arousal as it drives sex and
aggression in animal and human brains. Neurobiol Aging 24(Suppl 1):S83–S88. discussion
S91-2
Moore TR, Iams JD, Creasy RK, Burau KD, Davidson AL (1994) Diurnal and gestational patterns
of uterine activity in normal human pregnancy. The Uterine Activity in Pregnancy Working
Group. Obstet Gynecol 83(4):517–523
Moore RY, Speh JC, Leak RK (2002) Suprachiasmatic nucleus organization. Cell Tissue Res 309
(1):89–98. https://doi.org/10.1007/s00441-002-0575-2
Moore AM, Lucas KA, Goodman RL, Coolen LM, Lehman MN (2018) Three-dimensional
imaging of KNDy neurons in the mammalian brain using optical tissue clearing and multiple-
label immunocytochemistry. Sci Rep 8(1):2242. https://doi.org/10.1038/s41598-018-20563-2
Mounien L, Bizet P, Boutelet I, Gourcerol G, Fournier A, Vaudry H, Jegou S (2006) Pituitary
adenylate cyclase-activating polypeptide directly modulates the activity of proopiomelanocortin
neurons in the rat arcuate nucleus. Neuroscience 143(1):155–163. https://doi.org/10.1016/j.
neuroscience.2006.07.022
Muir AI, Chamberlain L, Elshourbagy NA, Michalovich D, Moore DJ, Calamari A, Szekeres PG,
Sarau HM, Chambers JK, Murdock P, Steplewski K, Shabon U, Miller JE, Middleton SE,
Darker JG, Larminie CG, Wilson S, Bergsma DJ, Emson P, Faull R, Philpott KL, Harrison DC
(2001) AXOR12, a novel human G protein-coupled receptor, activated by the peptide KiSS-1. J
Biol Chem 276(31):28969–28975. https://doi.org/10.1074/jbc.M102743200
440 N. Gotlieb et al.
Pan JT, Gala RR (1985) Central nervous system regions involved in the estrogen-induced afternoon
prolactin surge. II. Implantation studies. Endocrinology 117(1):388–395. https://doi.org/10.
1210/endo-117-1-388
Panda S, Sato TK, Castrucci AM, Rollag MD, DeGrip WJ, Hogenesch JB, Provencio I, Kay SA
(2002) Melanopsin (Opn4) requirement for normal light-induced circadian phase shifting.
Science 298(5601):2213–2216. https://doi.org/10.1126/science.1076848
Pau KY, Berria M, Hess DL, Spies HG (1993) Preovulatory gonadotropin-releasing hormone surge
in ovarian-intact rhesus macaques. Endocrinology 133(4):1650–1656
Piet R, Fraissenon A, Boehm U, Herbison AE (2015) Estrogen permits vasopressin signaling in
preoptic kisspeptin neurons in the female mouse. J Neurosci 35(17):6881–6892. https://doi.org/
10.1523/JNEUROSCI.4587-14.2015
Piet R, Dunckley H, Lee K, Herbison AE (2016) Vasoactive intestinal peptide excites GnRH
neurons in male and female mice. Endocrinology 157(9):3621–3630. https://doi.org/10.1210/
en.2016-1399
Pilorz V, Steinlechner S (2008) Low reproductive success in Per1 and Per2 mutant mouse females due
to accelerated ageing? Reproduction 135(4):559–568. https://doi.org/10.1530/REP-07-0434
Poletini MO, McKee DT, Kennett JE, Doster J, Freeman ME (2007) Knockdown of clock genes in
the suprachiasmatic nucleus blocks prolactin surges and alters FRA expression in the locus
coeruleus of female rats. Am J Phys Endocrinol Metab 293(5):E1325–E1334. https://doi.org/10.
1152/ajpendo.00341.2007
Poletini MO, Kennett JE, McKee DT, Freeman ME (2010) Central clock regulates the cervically
stimulated prolactin surges by modulation of dopamine and vasoactive intestinal polypeptide
release in ovariectomized rats. Neuroendocrinology 91(2):179–188. https://doi.org/10.1159/
000254379
Poling MC, Kim J, Dhamija S, Kauffman AS (2012) Development, sex steroid regulation, and
phenotypic characterization of RFamide-related peptide (Rfrp) gene expression and RFamide
receptors in the mouse hypothalamus. Endocrinology 153(4):1827–1840. https://doi.org/10.
1210/en.2011-2049
Rance NE (2009) Menopause and the human hypothalamus: evidence for the role of kisspeptin/
neurokinin B neurons in the regulation of estrogen negative feedback. Peptides 30(1):111–122.
https://doi.org/10.1016/j.peptides.2008.05.016
Ratajczak CK, Boehle KL, Muglia LJ (2009) Impaired steroidogenesis and implantation failure in
Bmal1-/- mice. Endocrinology 150(4):1879–1885. https://doi.org/10.1210/en.2008-1021
Ratajczak CK, Herzog ED, Muglia LJ (2010) Clock gene expression in gravid uterus and extra-
embryonic tissues during late gestation in the mouse. Reprod Fertil Dev 22(5):743–750. https://
doi.org/10.1071/RD09243
Ratajczak CK, Asada M, Allen GC, McMahon DG, Muglia LM, Smith D, Bhattacharyya S, Muglia
LJ (2012) Generation of myometrium-specific Bmal1 knockout mice for parturition analysis.
Reprod Fertil Dev 24(5):759–767. https://doi.org/10.1071/RD11164
Reppert SM, Henshaw D, Schwartz WJ, Weaver DR (1987) The circadian-gated timing of birth
in rats: disruption by maternal SCN lesions or by removal of the fetal brain. Brain Res
403(2):398–402
Resuehr D, Wildemann U, Sikes H, Olcese J (2007) E-box regulation of gonadotropin-releasing
hormone (GnRH) receptor expression in immortalized gonadotrope cells. Mol Cell Endocrinol
278(1–2):36–43. https://doi.org/10.1016/j.mce.2007.08.008
Rivkees SA, Reppert SM (1991) Appearance of melatonin receptors during embryonic life in
Siberian hamsters (Phodopus sungorous). Brain Res 568(1–2):345–349
Rizzo P, Raffone E, Benedetto V (2010) Effect of the treatment with myo-inositol plus folic acid
plus melatonin in comparison with a treatment with myo-inositol plus folic acid on oocyte
quality and pregnancy outcome in IVF cycles. A prospective, clinical trial. Eur Rev Med
Pharmacol Sci 14(6):555–561
Robertson MC, Friesen HG (1981) Two forms of rat placental lactogen revealed by radioimmuno-
assay. Endocrinology 108(6):2388–2390. https://doi.org/10.1210/endo-108-6-2388
442 N. Gotlieb et al.
Ronnekleiv OK, Kelly MJ (1988) Plasma prolactin and luteinizing hormone profiles during the
estrous cycle of the female rat: effects of surgically induced persistent estrus. Neuroendocrinol-
ogy 47(2):133–141. https://doi.org/10.1159/000124903
Ronnekleiv OK, Fang Y, Zhang C, Nestor CC, Mao P, Kelly MJ (2014) Research resource: gene
profiling of G protein-coupled receptors in the arcuate nucleus of the female. Mol Endocrinol 28
(8):1362–1380. https://doi.org/10.1210/me.2014-1103
Ruka KA, Burger LL, Moenter SM (2013) Regulation of arcuate neurons coexpressing kisspeptin,
neurokinin B, and dynorphin by modulators of neurokinin 3 and kappa-opioid receptors in adult
male mice. Endocrinology 154(8):2761–2771. https://doi.org/10.1210/en.2013-1268
Russo KA, La JL, Stephens SB, Poling MC, Padgaonkar NA, Jennings KJ, Piekarski DJ, Kauffman
AS, Kriegsfeld LJ (2015) Circadian control of the female reproductive axis through gated
responsiveness of the RFRP-3 system to VIP signaling. Endocrinology 156(7):2608–2618.
https://doi.org/10.1210/en.2014-1762
Saedi S, Khoradmehr A, Mohammad Reza JS, Tamadon A (2018) The role of neuropeptides and
neurotransmitters on kisspeptin/kiss1r-signaling in female reproduction. J Chem Neuroanat
92:71–82. https://doi.org/10.1016/j.jchemneu.2018.07.001
Sampaio RV, Conceicao S, Miranda MS, Sampaio Lde F, Ohashi OM (2012) MT3 melatonin
binding site, MT1 and MT2 melatonin receptors are present in oocyte, but only MT1 is present
in bovine blastocyst produced in vitro. Reprod Biol Endocrinol 10:103. https://doi.org/10.1186/
1477-7827-10-103
Samson WK, Burton KP, Reeves JP, McCann SM (1981) Vasoactive intestinal peptide stimulates
luteinizing hormone-releasing hormone release from median eminence synaptosomes. Regul
Pept 2(4):253–264
Sarkar DK, Chiappa SA, Fink G, Sherwood NM (1976) Gonadotropin-releasing hormone surge in
pro-oestrous rats. Nature 264(5585):461–463
Satake H, Hisada M, Kawada T, Minakata H, Ukena K, Tsutsui K (2001) Characterization of a
cDNA encoding a novel avian hypothalamic neuropeptide exerting an inhibitory effect on
gonadotropin release. Biochem J 354(Pt 2):379–385
Sawai N, Iijima N, Takumi K, Matsumoto K, Ozawa H (2012) Immunofluorescent histochemical
and ultrastructural studies on the innervation of kisspeptin/neurokinin B neurons to
tuberoinfundibular dopaminergic neurons in the arcuate nucleus of rats. Neurosci Res 74
(1):10–16. https://doi.org/10.1016/j.neures.2012.05.011
Sawai N, Iijima N, Ozawa H, Matsuzaki T (2014) Neurokinin B- and kisspeptin-positive fibers as
well as tuberoinfundibular dopaminergic neurons directly innervate periventricular hypophyseal
dopaminergic neurons in rats and mice. Neurosci Res 84:10–18. https://doi.org/10.1016/j.
neures.2014.05.002
Schafer D, Kane G, Colledge WH, Piet R, Herbison AE (2018) Sex- and sub region-dependent
modulation of arcuate kisspeptin neurons by vasopressin and vasoactive intestinal peptide. J
Neuroendocrinol:e12660. https://doi.org/10.1111/jne.12660
Schenker S, Yang Y, Perez A, Acuff RV, Papas AM, Henderson G, Lee MP (1998) Antioxidant
transport by the human placenta. Clin Nutr 17(4):159–167
Schoeller EL, Clark DD, Dey S, Cao NV, Semaan SJ, Chao LW, Kauffman AS, Stowers L, Mellon
PL (2016) Bmal1 is required for normal reproductive behaviors in male mice. Endocrinology
157(12):4914–4929. https://doi.org/10.1210/en.2016-1620
Schwartz WJ, Coleman RJ, Reppert SM (1983) A daily vasopressin rhythm in rat cerebrospinal
fluid. Brain Res 263(1):105–112
Seminara SB, Messager S, Chatzidaki EE, Thresher RR, Acierno JS Jr, Shagoury JK, Bo-Abbas Y,
Kuohung W, Schwinof KM, Hendrick AG, Zahn D, Dixon J, Kaiser UB, Slaugenhaupt SA,
Gusella JF, O’Rahilly S, Carlton MB, Crowley WF Jr, Aparicio SA, Colledge WH (2003) The
GPR54 gene as a regulator of puberty. N Engl J Med 349(17):1614–1627. https://doi.org/10.
1056/NEJMoa035322
Seron-Ferre M, Ducsay CA, Valenzuela GJ (1993) Circadian rhythms during pregnancy. Endocr
Rev 14(5):594–609. https://doi.org/10.1210/edrv-14-5-594
16 Development and Modulation of Female Reproductive Function by Circadian. . . 443
Szawka RE, Ribeiro AB, Leite CM, Helena CV, Franci CR, Anderson GM, Hoffman GE, Anselmo-
Franci JA (2010) Kisspeptin regulates prolactin release through hypothalamic dopaminergic
neurons. Endocrinology 151(7):3247–3257. https://doi.org/10.1210/en.2009-1414
Takahashi JS (2015) Molecular components of the circadian clock in mammals. Diabetes Obes
Metab 17(Suppl 1):6–11. https://doi.org/10.1111/dom.12514
Takayama H, Nakamura Y, Tamura H, Yamagata Y, Harada A, Nakata M, Sugino N, Kato H
(2003) Pineal gland (melatonin) affects the parturition time, but not luteal function and fetal
growth, in pregnant rats. Endocr J 50(1):37–43
Tamura H, Nakamura Y, Terron MP, Flores LJ, Manchester LC, Tan DX, Sugino N, Reiter RJ
(2008) Melatonin and pregnancy in the human. Reprod Toxicol 25(3):291–303. https://doi.org/
10.1016/j.reprotox.2008.03.005
Teclemariam-Mesbah R, Kalsbeek A, Pevet P, Buijs RM (1997) Direct vasoactive intestinal
polypeptide-containing projection from the suprachiasmatic nucleus to spinal projecting hypo-
thalamic paraventricular neurons. Brain Res 748(1–2):71–76
Thomas L, Purvis CC, Drew JE, Abramovich DR, Williams LM (2002) Melatonin receptors in
human fetal brain: 2-[(125)I]iodomelatonin binding and MT1 gene expression. J Pineal Res 33
(4):218–224
Thompson EL, Patterson M, Murphy KG, Smith KL, Dhillo WS, Todd JF, Ghatei MA, Bloom SR
(2004) Central and peripheral administration of kisspeptin-10 stimulates the hypothalamic-
pituitary-gonadal axis. J Neuroendocrinol 16(10):850–858. https://doi.org/10.1111/j.1365-
2826.2004.01240.x
Tian XZ, Wen Q, Shi JM, Liang W, Zeng SM, Tian JH, Zhou GB, Zhu SE, Liu GS (2010) Effects of
melatonin on in vitro development of mouse two-cell embryos cultured in HTF medium. Endocr
Res 35(1):17–23. https://doi.org/10.3109/07435800903539607
Topaloglu AK, Reimann F, Guclu M, Yalin AS, Kotan LD, Porter KM, Serin A, Mungan NO, Cook
JR, Imamoglu S, Akalin NS, Yuksel B, O’Rahilly S, Semple RK (2009) TAC3 and TACR3
mutations in familial hypogonadotropic hypogonadism reveal a key role for Neurokinin B in the
central control of reproduction. Nat Genet 41(3):354–358. https://doi.org/10.1038/ng.306
Torres-Farfan C, Rocco V, Monso C, Valenzuela FJ, Campino C, Germain A, Torrealba F,
Valenzuela GJ, Seron-Ferre M (2006) Maternal melatonin effects on clock gene expression in
a nonhuman primate fetus. Endocrinology 147(10):4618–4626. https://doi.org/10.1210/en.
2006-0628
Torres-Farfan C, Valenzuela FJ, Mondaca M, Valenzuela GJ, Krause B, Herrera EA, Riquelme R,
Llanos AJ, Seron-Ferre M (2008) Evidence of a role for melatonin in fetal sheep physiology:
direct actions of melatonin on fetal cerebral artery, brown adipose tissue and adrenal gland. J
Physiol 586(16):4017–4027. https://doi.org/10.1113/jphysiol.2008.154351
Trevisan CM, Montagna E, de Oliveira R, Christofolini DM, Barbosa CP, Crandall KA, Bianco B
(2018) Kisspeptin/GPR54 system: what do we know about its role in human reproduction? Cell
Physiol Biochem 49(4):1259–1276. https://doi.org/10.1159/000493406
Tsutsui K, Saigoh E, Ukena K, Teranishi H, Fujisawa Y, Kikuchi M, Ishii S, Sharp PJ (2000) A
novel avian hypothalamic peptide inhibiting gonadotropin release. Biochem Biophys Res
Commun 275(2):661–667. https://doi.org/10.1006/bbrc.2000.3350
Tsutsui K, Saigoh E, Yin H, Ubuka T, Chowdhury VS, Osugi T, Ukena K, Sharp PJ, Wingfield JC,
Bentley GE (2009) A new key neurohormone controlling reproduction, gonadotrophin-
inhibitory hormone in birds: discovery, progress and prospects. J Neuroendocrinol 21
(4):271–275. https://doi.org/10.1111/j.1365-2826.2009.01829.x
Ubuka T, Tsutsui K (2018) Comparative and evolutionary aspects of gonadotropin-inhibitory
hormone and FMRFamide-like peptide systems. Front Neurosci 12:747. https://doi.org/10.
3389/fnins.2018.00747
Ubuka T, Inoue K, Fukuda Y, Mizuno T, Ukena K, Kriegsfeld LJ, Tsutsui K (2012) Identification,
expression, and physiological functions of Siberian hamster gonadotropin-inhibitory hormone.
Endocrinology 153(1):373–385. https://doi.org/10.1210/en.2011-1110
16 Development and Modulation of Female Reproductive Function by Circadian. . . 445
Unfer V, Raffone E, Rizzo P, Buffo S (2011) Effect of a supplementation with myo-inositol plus
melatonin on oocyte quality in women who failed to conceive in previous in vitro fertilization
cycles for poor oocyte quality: a prospective, longitudinal, cohort study. Gynecol Endocrinol 27
(11):857–861. https://doi.org/10.3109/09513590.2011.564687
van der Beek EM, van Oudheusden HJ, Buijs RM, van der Donk HA, van den Hurk R, Wiegant VM
(1994) Preferential induction of c-fos immunoreactivity in vasoactive intestinal polypeptide-
innervated gonadotropin-releasing hormone neurons during a steroid-induced luteinizing hor-
mone surge in the female rat. Endocrinology 134(6):2636–2644. https://doi.org/10.1210/endo.
134.6.8194489
Van der Beek EM, Horvath TL, Wiegant VM, Van den Hurk R, Buijs RM (1997) Evidence for a
direct neuronal pathway from the suprachiasmatic nucleus to the gonadotropin-releasing hor-
mone system: combined tracing and light and electron microscopic immunocytochemical
studies. J Comp Neurol 384(4):569–579
van der Horst GT, Muijtjens M, Kobayashi K, Takano R, Kanno S, Takao M, de Wit J, Verkerk A,
Eker AP, van Leenen D, Buijs R, Bootsma D, Hoeijmakers JH, Yasui A (1999) Mammalian
Cry1 and Cry2 are essential for maintenance of circadian rhythms. Nature 398(6728):627–630.
https://doi.org/10.1038/19323
Vanecek J (1988) The melatonin receptors in rat ontogenesis. Neuroendocrinology 48(2):201–203.
https://doi.org/10.1159/000125008
Vatish M, Steer PJ, Blanks AM, Hon M, Thornton S (2010) Diurnal variation is lost in preterm
deliveries before 28 weeks of gestation. BJOG 117(6):765–767. https://doi.org/10.1111/j.1471-
0528.2010.02526.x
Vida B, Deli L, Hrabovszky E, Kalamatianos T, Caraty A, Coen CW, Liposits Z, Kallo I (2010)
Evidence for suprachiasmatic vasopressin neurones innervating kisspeptin neurones in the
rostral periventricular area of the mouse brain: regulation by oestrogen. J Neuroendocrinol 22
(9):1032–1039. https://doi.org/10.1111/j.1365-2826.2010.02045.x
Vilches N, Spichiger C, Mendez N, Abarzua-Catalan L, Galdames HA, Hazlerigg DG, Richter HG,
Torres-Farfan C (2014) Gestational chronodisruption impairs hippocampal expression of
NMDA receptor subunits Grin1b/Grin3a and spatial memory in the adult offspring. PLoS
One 9(3):e91313. https://doi.org/10.1371/journal.pone.0091313
Waddell BJ, Wharfe MD, Crew RC, Mark PJ (2012) A rhythmic placenta? Circadian variation,
clock genes and placental function. Placenta 33(7):533–539. https://doi.org/10.1016/j.placenta.
2012.03.008
Wakabayashi Y, Nakada T, Murata K, Ohkura S, Mogi K, Navarro VM, Clifton DK, Mori Y,
Tsukamura H, Maeda K, Steiner RA, Okamura H (2010) Neurokinin B and dynorphin A in
kisspeptin neurons of the arcuate nucleus participate in generation of periodic oscillation of
neural activity driving pulsatile gonadotropin-releasing hormone secretion in the goat. J
Neurosci 30(8):3124–3132. https://doi.org/10.1523/JNEUROSCI.5848-09.2010
Wakabayashi Y, Yamamura T, Sakamoto K, Mori Y, Okamura H (2013) Electrophysiological and
morphological evidence for synchronized GnRH pulse generator activity among Kisspeptin/
neurokinin B/dynorphin a (KNDy) neurons in goats. J Reprod Dev 59(1):40–48
Wang Y, Gu F, Deng M, Guo L, Lu C, Zhou C, Chen S, Xu Y (2016) Rotating shift work and
menstrual characteristics in a cohort of Chinese nurses. BMC Womens Health 16(1):24. https://
doi.org/10.1186/s12905-016-0301-y
Welsh DK, Yoo SH, Liu AC, Takahashi JS, Kay SA (2004) Bioluminescence imaging of individual
fibroblasts reveals persistent, independently phased circadian rhythms of clock gene expression.
Curr Biol 14(24):2289–2295. https://doi.org/10.1016/j.cub.2004.11.057
Wharfe MD, Mark PJ, Waddell BJ (2011) Circadian variation in placental and hepatic clock genes
in rat pregnancy. Endocrinology 152(9):3552–3560. https://doi.org/10.1210/en.2011-0081
Wiegand SJ, Terasawa E (1982) Discrete lesions reveal functional heterogeneity of suprachiasmatic
structures in regulation of gonadotropin secretion in the female rat. Neuroendocrinology 34
(6):395–404. https://doi.org/10.1159/000123335
446 N. Gotlieb et al.
Wiegand SJ, Terasawa E, Bridson WE, Goy RW (1980) Effects of discrete lesions of preoptic and
suprachiasmatic structures in the female rat. Alterations in the feedback regulation of gonado-
tropin secretion. Neuroendocrinology 31(2):147–157. https://doi.org/10.1159/000123066
Williams WP 3rd, Kriegsfeld LJ (2012) Circadian control of neuroendocrine circuits regulating
female reproductive function. Front Endocrinol 3:60. https://doi.org/10.3389/fendo.2012.00060
Williams LM, Martinoli MG, Titchener LT, Pelletier G (1991) The ontogeny of central melatonin
binding sites in the rat. Endocrinology 128(4):2083–2090. https://doi.org/10.1210/endo-128-4-2083
Williams WP 3rd, Jarjisian SG, Mikkelsen JD, Kriegsfeld LJ (2011) Circadian control of kisspeptin
and a gated GnRH response mediate the preovulatory luteinizing hormone surge. Endocrinol-
ogy 152(2):595–606. https://doi.org/10.1210/en.2010-0943
Wintermantel TM, Elzer J, Herbison AE, Fritzemeier KH, Schutz G (2006) Genetic dissection of
estrogen receptor signaling in vivo. Ernst Schering Found Symp Proc 1:25–44
Wu M, Dumalska I, Morozova E, van den Pol AN, Alreja M (2009) Gonadotropin inhibitory
hormone inhibits basal forebrain vGluT2-gonadotropin-releasing hormone neurons via a direct
postsynaptic mechanism. J Physiol 587(Pt 7):1401–1411. https://doi.org/10.1113/jphysiol.
2008.166447
Xie M, Utzinger KS, Blickenstorfer K, Leeners B (2018) Diurnal and seasonal changes in semen
quality of men in subfertile partnerships. Chronobiol Int 35(10):1375–1384. https://doi.org/10.
1080/07420528.2018.1483942
Yamazaki S, Numano R, Abe M, Hida A, Takahashi R, Ueda M, Block GD, Sakaki Y, Menaker M,
Tei H (2000) Resetting central and peripheral circadian oscillators in transgenic rats. Science
288(5466):682–685
Yano T, Iijima N, Kakihara K, Hinuma S, Tanaka M, Ibata Y (2003) Localization and neuronal
response of RFamide related peptides in the rat central nervous system. Brain Res 982
(2):156–167
Yeo SH, Colledge WH (2018) The role of Kiss1 neurons as integrators of endocrine, metabolic, and
environmental factors in the hypothalamic-pituitary-gonadal axis. Front Endocrinol 9:188.
https://doi.org/10.3389/fendo.2018.00188
Yip SH, Boehm U, Herbison AE, Campbell RE (2015) Conditional viral tract tracing delineates the
projections of the distinct Kisspeptin neuron populations to gonadotropin-releasing hormone
(GnRH) neurons in the mouse. Endocrinology 156(7):2582–2594. https://doi.org/10.1210/en.
2015-1131
Yogev L, Terkel J (1980) Effects of photoperiod, absence of photic cues, and suprachiasmatic
nucleus lesions on nocturnal prolactin surges in pregnant and pseudopregnant rats. Neuroendo-
crinology 31(1):26–33. https://doi.org/10.1159/000123046
Young J, Bouligand J, Francou B, Raffin-Sanson ML, Gaillez S, Jeanpierre M, Grynberg M,
Kamenicky P, Chanson P, Brailly-Tabard S, Guiochon-Mantel A (2010) TAC3 and TACR3
defects cause hypothalamic congenital hypogonadotropic hypogonadism in humans. J Clin
Endocrinol Metab 95(5):2287–2295. https://doi.org/10.1210/jc.2009-2600
Zahn V, Hattensperger W (1993) Circadian rhythm of pregnancy contractions. Z Geburtshilfe
Perinatol 197(1):1–10
Zhao S, Kriegsfeld LJ (2009) Daily changes in GT1-7 cell sensitivity to GnRH secretagogues that
trigger ovulation. Neuroendocrinology 89(4):448–457. https://doi.org/10.1159/000192370
Zhao L, Zhong M, Xue HL, Ding JS, Wang S, Xu JH, Chen L, Xu LX (2014) Effect of RFRP-3 on
reproduction is sex- and developmental status-dependent in the striped hamster (Cricetulus
barabensis). Gene 547(2):273–279. https://doi.org/10.1016/j.gene.2014.06.054
Glossary
Caudal brainstem Including the medulla (medulla oblongata), aka the hindbrain.
Houses the area postrema (AP), nucleus of the solitary tract (NTS), dorsal motor
nucleus of the vagus (DMV), medullary reticular formation (RF), and ventrolat-
eral medulla (VLM).
Cell lineage All the cells born from one dividing cell (or from one specific kind of
dividing cell) form a “cell lineage.”
Cellular central oscillator Activity of the core circadian oscillator at the cellular
(as opposed to organ or organismal) level.
CHIP-seq High-throughput sequencing method combining chromatin immunopre-
cipitation technique with deep sequencing. It allows the identification of specific
DNA sequences bound to proteins of interest in vivo. Benefits include single
nucleotide resolution, relatively low amount of DNA input, and high dynamic
range for quantification.
Chronobiology The study of endogenous biological rhythms and the entrainment
by environmental cues. Also referred to as biochronometry.
Chronotype Intrinsic, individual differences in the synchrony between light cycles
and activity rhythms. Manifests as “morningness” (phase-advanced) and
“eveningness” (phase-delayed sleep/wake rhythms).
Circadian An endogenous behavioural, biochemical, or physiological oscillation
that approximates a 24-h period.
Circadian disruption An incongruence between endogenous circadian rhythms
and external time.
Circadian entrainment This occurs when rhythmic physiological or behavioral
events match their period to that of an environmental oscillation. Entraining
signals can include light, food, and neurotransmitters.
Circadian phase synchrony and stabilization Circadian phase synchrony occurs
when cellular clocks are synchronized and run in phase, typically in response to
an external cue. Circadian phase stabilization indicates that this synchrony is
robust and reproducible over many phases.
Circadian rhythms Endogenously generated behavioral, physiological, endocrine,
metabolic, and/or molecular rhythms having a ~24-h period that is synchronized
to environmental time and persist in the absence of environmental cues, but can
be entrained by light, food and other external cues.
Clock gene A gene that is critical for generating circadian timekeeping at the
cellular level.
Collateral inhibition Neural networks in which inhibitory (usually GABAergic)
neurons inhibit the activity of synaptically connected neurons.
Conditional mutant A transgenic mouse mutant line engineered so that gene A
(gene of interest) will become inactive only in a certain specific tissue domain
where it is co-expressed with gene B (“Cre-driver gene”).
Core circadian oscillator A transcriptional feedback loop in which increased
expression of Per and Cry proteins inhibits activity of Bmal1 and Clock. It is
active in virtually all mammalian cells.
Glossary 449
Emetic reflex The vomiting reflex, present in most mammalian species (including
humans), but absent in non-emetic rodents (e.g., rats and mice).
Endocannabinoids A class of signaling molecules derived from membrane lipids
and that bind to the same receptors as the active compounds in marijuana.
Endogenous Naturally occurring, e.g., hormones released from endocrine tissues.
Entrainment Synchronization of endogenously generated circadian rhythms to the
external environment.
Epigenetic modification Biochemical change to the genome or chromatin template
that alters the conformational structure and induces either a permissive or an
inhibitory state for gene transcription. Examples include DNA methylation,
histone acetylation and histone methylation.
Epigenetic regulator Chromatin remodeling enzymes and enzymes that place or
remove methylation and acetylation marks on histones bound to chromatin,
altering access of transcription factors to bind DNA.
Epigenetics It refers to modifications to the DNA that impact gene expression but
do not change the genetic code (i.e., G, C, A & T). These can include direct
methylation of the DNA and indirect via modifications of the surrounding
histones that impact accessibility of transcription factors to the DNA.
Epithelial to mesenchymal-like transition Epithelial cells lose their polarity,
release cell–cell adhesion, change shape and acquire migratory properties, a
process that occurs in development of many organs and in cancer.
Estradiol A sex steroid of the estrogen family, synthesized in, and secreted from,
the ovary.
Estrogens A class of steroid hormones that includes estradiol, estriol, and estrone.
In adulthood estradiol is produced in high quantities by the ovary. During
development estradiol is produced in the brain by being derived from testosterone
in a process called aromatization. In rodents, estrogens are masculinizing
hormones during development.
Exogenous Artificially sourced, e.g., drugs, synthetic peptides, and other chemical
agents.
Exteroceptive Sensory signals arising from the environment (e.g., stimuli con-
veyed to the central nervous system through visual or auditory pathways) rather
than from within the body.
Fast neurotransmitter Fast-acting neurotransmitters such as glutamate (excit-
atory) and gamma-aminobutyric acid (GABA) (inhibitory), modulate their target
cells in under one millisecond by directly activating ligand-gated ion channels.
Monoamines and neuropeptides, which act via G-protein coupled receptors
(GPCRs), act more slowly.
Fate mapping Any technique allowing for the labeling of a cell lineage, i.e., the
labeling of one specific kind of dividing cell and all the cells that derive from it.
Floor plate A tissue domain, that acts of a source of morphogens, occupying the
ventral midline of the developing neural tube and formed by specific cells (not
neuroepithelial cells) with essential functions in neural tube fate specification.
The ventral midline of the most rostral part of the neural tube (corresponding to
Glossary 451
the hypothalamus) shows important differences with the floor plate elsewhere and
usually receives a different name (see Hypothalamic floor plate, Rostral
diencephalic ventral midline).
Follicle stimulating hormone A hormone secreted by the anterior pituitary that
stimulates the growth of ovarian follicles and sperm maturation.
Folliculostellate cells Non-endocrine cells in the anterior pituitary with star-like
cytoplasmic processes.
Forebrain Consists of structures derived from the telencephalon and diencepha-
lon. It is rostral to the mesencephalon and contains the centers involved in the
control of sensorimotor, autonomic and endocrine functions, in emotional behav-
ior, as well as in cognitive functions such as learning and memory.
Fos The protein product of the immediate-early gene, c-fos. Fos protein accumulates
within the nuclei of neurons and other cells due to increased intracellular calcium.
Increased neuronal Fos labeling (compared to baseline levels) is interpreted as
evidence that the neuron was stimulated/activated (e.g., received increased excit-
atory synaptic input).
Free run A self-sustained circadian rhythm that is not synchronized to
environmental time.
GABA Gamma-aminobutyric acid—The primary inhibitory neurotransmitter in the
adult brain and a secreted factor that regulates cell function during development.
GAD Glutamic acid decarboxylase—Enzyme that converts glutamate to GABA
(two forms GAD65 and GAD67).
GATA2 Transcription factor that drives thyrotrope and gonadotrope cell fate.
Genomic imprinting Parent-of-origin-dependent, monoallelic gene expression
caused by epigenetic marks (i.e., DNA methylation) set in the parental germ
cells and maintained in the somatic cells of the offspring.
GHRH-neurons Arcuate nucleus neurons that stimulate linear growth; Kiss1-
neurons, arcuate nucleus neurons that directs reproduction.
Gliogenesis Generation of new glial cells from progenitor cells.
Glycoprotein hormones Bioactive LH, FSH, and TSH are composed of specific
beta subunits and a common alpha subunit, CGA or chorionic gonadotropin
alpha. Both alpha and beta subunits are glycosylated.
GnRH Gonadotropin-releasing hormone—A hormone secreted into anterior pitui-
tary portal system that stimulates the release of gonadotropins from the anterior
pituitary into the general circulation and as such regulates reproductive function.
Gonadotropes Anterior lobe cells that produce follicle-stimulating hormone (FSH)
and luteinizing hormone (LH).
Gonadotropin-inhibitory hormone An RFamide peptide that is inhibitory to the
synthesis and secretion of gonadotropins from the anterior pituitary; a ligand for
GPR147 also called RFamide-related peptide-3 (RFRP-3) in mammals.
Gonadotropins Hormones synthesized and secreted from the anterior pituitary that
regulate sex steroid production and gametogenesis—follicle stimulating hormone
and luteinizing hormone, respectively.
HESX1 Transcription factor involved in development of the eye, forebrain, and
Rathke’s pouch.
452 Glossary
Prechordal plate The portion of the axial mesoderm underlying the presumptive
hypothalamus. Shh from the prechordal plate is required for hypothalamic
specifications.
Preoptic Area (POA) A brain region critical to the control of male sexual behavior
and female maternal behavior. It is also the site of the sexually dimorphic nucleus
(SDN), and a variety of other sex differences in anatomy and physiology.
Primary wake-promoting neurons Neurons that rapidly (i.e., within seconds to
minutes) induce wakefulness following activation.
Progenitor cell A proliferative cell that can undergo a limited number of cell
divisions, and gives rise to a limited type of differentiated cells.
Progenitor domain A restricted region of the early neural tube, which generates
the cells for a certain subdivision of the central nervous system.
PROP1 Pituitary-specific transcription factor that marks anterior and intermediate
lobe progenitors and is required for expansion of the Pou1f1 lineage.
Prostaglandin E2 (PGE2) A membrane derived signaling molecule usually
associated with inflammation but which also has normal roles in brain develop-
ment, including the promoting of synapse formation in the male POA.
Protein kinase-A (PKA) An enzyme which phosphorylates other proteins to
modify their function, such as subunits of glutamate receptors, which impacts
how they traffic to the membrane.
Puberty Developmental period when an individual becomes capable of sexual
reproduction. It is marked by genital organ maturation and development of
secondary sexual characteristics.
PVN Paraventricular nucleus of the hypothalamus—Heterogeneous grouping of
cells in the hypothalamus known to function in the regulation of blood pressure,
appetite, stress responses, sexual function, cardiac output, and organismal
growth.
Rapid eye movement (REM) sleep Comprises ~20% of sleep time in young adult
humans. During REM sleep, brain waves as measured by electroencephalo-
graphic recording are typically high frequency and low amplitude, and the sleeper
exhibits flaccid paralysis, with exception of the ocular muscles, which move
rapidly. Dreaming is frequent and vivid.
Rathke’s pouch An invagination of oral ectoderm that forms the anterior and
intermediate lobes of the pituitary gland.
Retinohypothalamic tract Axonal projections of the SCN to the hypothalamus.
RFamide peptide A peptide that contains Arg-Phe-NH2 motif at its C-terminus.
RFRP-3 An RFamide peptide synthesized and secreted from the hypothalamus that
inhibits the release of gonadotropins from the anterior pituitary; mammalian
ortholog of gonadotropin-inhibitory hormone (GnIH).
Rhythmic epigenetics The study of the biochemical, physiological and environ-
mental regulation of epigenetic enzyme oscillations and the functional signifi-
cance of predictable changes in epigenetic modifications across different
biological periods (e.g. seasonal).
456 Glossary
VNO Vomeronasal organ—Putative site of origin for GnRH neurons and origina-
tion of olfactory fibers to the accessory olfactory bulb.
ZEB1 Transcription factor implicated in differentiation of lactotropes from
somatotropes.
ZEB2 Transcription factor involved in epithelial to mesenchymal-like transition of
pituitary progenitors.
Zona incerta (ZI) A bilateral elongated nucleus in dorsal hypothalamus. It
regulates a diverse range of physiological processes, including encoding sleep
pressure.
Index
E F
EAP1, 209 FastQC, 217
Earlier puberty, 210 Fate-mapping, 37
462 Index