ArticlePDF AvailableLiterature Review

Peripheral and central mechanisms involved in hormonal control of male and female reproduction

Authors:

Abstract and Figures

Reproduction involves the integration of hormonal signals acting across multiple systems to generate a synchronized physiological output. A critical component of reproduction is the luteinizing hormone (LH) surge, which is mediated by estradiol (E2) and neuroprogesterone interacting to stimulate kisspeptin release in the rostral periventricular nucleus of the third ventricle in rats. Recent evidence has shown that both classical and membrane E2 and progesterone signaling is involved in this pathway. A metabolite of gonadotropin-releasing hormone (GnRH), GnRH-(1-5), has been shown to stimulate GnRH expression, secretion, and has a role in the regulation of lordosis. Additionally, gonadotropin-inhibitory hormone (GnIH) projects to and influences the activity of GnRH neurons in birds. Stress-induced changes in GnIH have been shown to alter breeding behaviors in birds, demonstrating another molecular control of reproduction. Peripherally, paracrine and autocrine actions within the gonad have been suggested as therapeutic targets for infertility in both males and females. Dysfunction of testicular prostaglandin synthesis is a possible cause of idiopathic male infertility. Indeed, local production of melatonin and corticotropin-releasing hormone (CRH) could influence spermatogenesis via immune pathways in the gonad. In females, vascular endothelial growth factor A (VEGF-A) has been implicated in an angiogenic process that mediates development of the corpus luteum and thus fertility via the Notch signaling pathway. Age-induced decreases in fertility involve ovarian kisspeptin and its regulation of ovarian sympathetic innervation. Finally, morphological changes in the arcuate nucleus of the hypothalamus influence female sexual receptivity in rats. The processes mediating these morphological changes have been shown to involve rapid effects of E2 controlling synaptogenesis in this hypothalamic nucleus. Together, this review highlights new research in these areas, focusing on recent findings in the molecular mechanisms of central and peripheral hormonal control of reproduction. This article is protected by copyright. All rights reserved.
Content may be subject to copyright.
Journal of Neuroendocrinology, 2016, 28, 10.1111/jne.12405
REVIEW ARTICLE ©2016 British Society for Neuroendocrinology
Peripheral and Central Mechanisms Involved in the Hormonal Control of
Male and Female Reproduction
L. M. Rudolph*, G. E. Bentley, R. S. Calandra, A. H. Paredes§, M. Tesone,T.J.Wuand P. E. Micevych*
*Department of Neurobiology, Laboratory of Neuroendocrinology, David Geffen School of Medicine at UCLA, Los Angeles, CA, USA.
Department of Integrative Biology, and Helen Wills Neuroscience Institute, University of California, Berkeley, Berkeley, CA, USA.
Instituto de Biolog
ıa y Medicina Experimental (IBYME-CONICET), Buenos Aires, Argentina.
§Laboratory of Neurobiochemistry, Faculty of Chemistry and Pharmaceutical Sciences, Universidad de Chile, Independencia, Santiago, Chile.
Department of Obstetrics and Gynecology, Center for Neuroscience and Regenerative Medicine, Uniformed Services University, Bethesda, MD, USA.
Journal of
Neuroendocrinology
Correspondence to: Lauren M.
Rudolph, UCLA Department of
Neurobiology, CHS 73-074, 650
Charles E. Young Dr. S., Los Angeles,
CA 90095, USA (e-mail:
lrudolph@mednet.ucla.edu).
Reproduction involves the integration of hormonal signals acting across multiple systems to
generate a synchronised physiological output. A critical component of reproduction is the lutein-
ising hormone (LH) surge, which is mediated by oestradiol (E
2
) and neuroprogesterone interact-
ing to stimulate kisspeptin release in the rostral periventricular nucleus of the third ventricle in
rats. Recent evidence indicates the involvement of both classical and membrane E
2
and proges-
terone signalling in this pathway. A metabolite of gonadotrophin-releasing hormone (GnRH),
GnRH-(1-5), has been shown to stimulate GnRH expression and secretion, and has a role in the
regulation of lordosis. Additionally, gonadotrophin release-inhibitory hormone (GnIH) projects to
and influences the activity of GnRH neurones in birds. Stress-induced changes in GnIH have
been shown to alter breeding behaviour in birds, demonstrating another mechanism for the
molecular control of reproduction. Peripherally, paracrine and autocrine actions within the
gonad have been suggested as therapeutic targets for infertility in both males and females. Dys-
function of testicular prostaglandin synthesis is a possible cause of idiopathic male infertility.
Indeed, local production of melatonin and corticotrophin-releasing hormone could influence
spermatogenesis via immune pathways in the gonad. In females, vascular endothelial growth
factor A has been implicated in an angiogenic process that mediates development of the corpus
luteum and thus fertility via the Notch signalling pathway. Age-induced decreases in fertility
involve ovarian kisspeptin and its regulation of ovarian sympathetic innervation. Finally, mor-
phological changes in the arcuate nucleus of the hypothalamus influence female sexual recep-
tivity in rats. The processes mediating these morphological changes have been shown to involve
the rapid effects of E
2
controlling synaptogenesis in this hypothalamic nucleus. In summary, this
review highlights new research in these areas, focusing on recent findings concerning the
molecular mechanisms involved in the central and peripheral hormonal control of reproduction.
Key words: progesterone, oestrogens, androgens, paracrine, autocrine
doi: 10.1111/jne.12405
Introduction
Reproduction is tightly regulated by the actions of hormones, both
central and peripheral in origin. The ‘classical’ mechanisms of ster-
oidal control of reproduction have been studied for decades, yet
questions remain about how these hormones interact within the
nervous system to elicit a coordinated response leading to ovula-
tion and fertilisation. The common final pathway to the regulation
of reproductive function is dependent on the appropriate
functioning of the hypothalamic-pituitary-gonadal (HPG) axis. The
proper coordination of the HPG axis relies largely on the inputs
that regulate gonadotrophin-releasing hormone (GnRH) release
from hypothalamic neurones. In recent years, numerous nonclassi-
cal mechanisms have been uncovered, including newly understood
membrane, autocrine and paracrine actions of steroid hormones. In
addition, novel neuropeptides have been added to the list of
neuroendocrine mediators such as the truncated GnRH [GnRH-(1-
5)], as well as the inhibitory gonadotrophin release-inhibitory hor-
mone (GnIH). Together, these recently appreciated events have
changed our understanding of the interaction of the HPG axis and
the relationship between the periphery and the central nervous sys-
tem in the regulation of reproduction.
Control of the LH surge
Central nervous system (CNS) regulation of the LH surge
As reviewed previously, oestradiol membrane signalling, comprising
oestradiol (E
2
) signalling that is initiated at the cell membrane,
plays an important role in the CNS synthesis of progesterone (neu-
roP) needed for oestrogen positive-feedback of the LH surge (1).
Although the preovulatory rise in circulating E
2
is essential for
stimulating gonadotrophin release (2–4), progesterone is also nec-
essary for the LH surge (5–9). In ovariectomised rats and mice, E
2
induces an LH release (10) and LH levels are augmented by addi-
tional application of progesterone (11,12). Blocking progesterone
receptor (PR) or progesterone synthesis prevents the E
2
-induced
GnRH and LH surges in ovariectomised rats (5,13) and arrests the
oestrous cycle in intact female rats (14). Most critically for this dis-
cussion, ablation of PR in kisspeptin (KP)-expressing neurones abro-
gates oestrogen positive-feedback (15), indicating that that both E
2
and progesterone are necessary for surge release of LH.
Where does neuroP act to influence the LH surge? It is well
established that GnRH neurones themselves do not express the req-
uisite steroid hormone receptors, oestrogen receptor (ER)aand PR
(16,17). There is now solid evidence that the LH surge ‘pattern gen-
erator’, which integrates steroid hormone information and regulates
oestrogen positive-feedback is a population of KP-expressing neu-
rones of the rostral periventricular nucleus of the third ventricle
(RP3V), an area that includes the anterior periventricular nucleus
and the anteroventral periventricular nucleus (18–25). Kiss1 neu-
rones in the RP3V are critical for GnRH secretion because KP
released from Kiss1 neurones activates GnRH neurones via GPR54,
a G-protein coupled receptor that binds KP (26–28). Although much
of the work on steroid regulation of KP and its gene, Kiss1, has
focused on E
2
(29,30), it is now evident that E
2
and neuroP func-
tion together to regulate KP. First, both ERaand PR are needed for
positive-feedback of the LH surge (31,32), and both have been loca-
lised in KP neurones, although neither are found in GnRH neurones
(20,33). Consistent with the need for E
2
-induced PRs for the LH
surge, a substantial number of KP neurones in RP3V and the
arcuate nucleus of the hypothalamus (ARH) express PR after E
2
treatment (25,30,33,34). Coincident with this, rising E
2
levels during
pro-oestrus induce neuroP synthesis (14,35).
A combination of in vitro and in vivo experiments have demon-
strated that neuroP acts on KP neurones to mediate oestrogen pos-
itive-feedback (Fig. 1). Integrated steroid signalling was studied in a
cell line (mHypoA51s) that approximates ‘sexually mature’ female
hypothalamic neurones. These immortalised neurones have the
characteristics of post-pubertal RP3V KP neurones because they
express ERa, PR and KP (36). As with KP neurones in vivo,E
2
and
the ERaagonist, PPT, induced KP and PR in mHypoA51s.
Significantly, E
2
-induced PR up-regulation was dependent on an
intracellular ER, whereas KP expression was stimulated by membrane-
impermeable E
2
(E
2
coupled to bovine serum albumin; E-6-BSA). These
data suggest that anterior hypothalamic KP neurones utilise both
membrane-initiated and classical nuclear oestrogen signalling to up-
regulate KP and PR, which are essential for the LH surge.
The nature of progesterone signalling in KP neurones remains to
be clarified. In addition to classical nuclear PR, there are intriguing
suggestions that KP neurones in vitro and in vivo have membrane
progesterone receptors, especially mPRb(37). The mPRs are seven-
transmembrane proteins that activate G proteins that belong to the
progestin and adipoQ receptor (PAQR) family not the classic G pro-
tein-coupled receptor (GPCR) family (38–40). PAQRs can signal
through mitogen-activated protein kinase activation and increasing
[Ca
2+
]
i
(41–47); but see also (48). Studies in mHypoA51s indicate
that classical PR is responsible for progesterone-induced signalling
events. Treatment of E
2
-primed mHypoA51s with progesterone
induces a rapid increase in free cytoplasmic calcium ([Ca
2+
]
i
), which
appears to be responsible for the release of KP induced by
progesterone, whereas inhibition with RU486 prevents the [Ca
2+
]
i
increase (36).
In vivo, preliminary experiments have demonstrated that exoge-
nous progesterone rescued the LH surge in females whose hypothala-
mic steroidogenesis was blocked with the CYP11A1 inhibitor
aminoglutethimide (AGT) (49). In AGT-treated animals, infusions of
progesterone or KP into the diagonal band of Broca induced an LH
surge, confirming that KP operates downstream of neuroP. Finally, KP
knockdown in the RP3V prevented the E
2
-induced LH surge (49). Most
importantly, the ablation of PR in KP neurones in ovariectomised mice
abrogates E
2
positive-feedback (15) demonstrating that that both E
2
and neuroP are necessary for the surge release of LH.
Molecular mechanisms of GnRH-(1-5) action
The decapeptide GnRH (pGlu-His-Trp-Ser-Tyr-Gly-Leu-Arg-Pro-Gly-
NH2) is highly conserved across species, suggesting its functional
importance throughout evolution (50). GnRH is primarily known for
its role in regulating reproductive function and behaviour via inter-
action with KP and its cognate receptor, GPR54, in the hypothala-
mus (51–57). Within each oestrous cycle, a rapid increase in GnRH
secretion culminates in an LH surge, which precedes the onset of
sexual receptivity and ovulation. In addition to its effects on the
secretion of LH, GnRH can autoregulate its own biosynthesis and
secretion via an ultrashort-loop feedback mechanism (58–62).
GnRH not only functions in its full form, but also can signal via
its metabolite, GnRH-(1-5). GnRH-(1-5) is produced by the cleavage
of GnRH by the zinc metalloendopeptidase EC3.4.24.15 (EP24.15) at
the covalent bond linking the fifth and sixth amino acids (63–65)
(Fig. 2). Localisation of EP24.15 supports the involvement of
EP24.15 in the modulation of hypothalamic GnRH neuronal func-
tion (63,66). EP24.15 immunoreactivity is sensitive to hormonal
fluctuations: increasing on pro-oestrous day of the rat oestrous
cycle within the median eminence, with a peak expression coincid-
ing with the LH surge (63). Unlike GnRH, GnRH-(1-5) robustly
©2016 British Society for Neuroendocrinology Journal of Neuroendocrinology, 2016, 28, 10.1111/jne.12405
2of12 L. M. Rudolph et al.
stimulates GnRH gene expression (67) and stimulates GnRH secre-
tion (68). Moreover, the GnRH facilitation of lordosis behaviour is
actually mediated by its metabolism to GnRH-(1-5) (69).
Interestingly, studies show that GnRH-(1-5) does not bind to
the GnRH receptor (51) but binds to two orphan GPCRs: GPR101
(70) and GPR173 (71,72) (Fig. 2). Both GPR101 and GPR173 are
members of the Rhodopsin class of receptors. The Rhodopsin fam-
ily is the largest of the five groups of orphan receptors with 672
members of which 63 have no known ligands. Both GnRH-(1-5)-
binding GPCRs are highly conserved and are highly expressed in
the hypothalamus (Allen Brain Bank) (73,74). In several species,
the coding sequence for GPR101 is located on the X chromosome
in a band that is syntenic between species (75). In mouse,
GPR101 mRNA is 2186 bases, encoding a seven-transmembrane
receptor that is approximately 51 kDa (76). The sequenced
GPR173 mRNA is 1122 bases, which translates to a 42-kDa
seven-transmembrane receptor (73). Functional studies suggest
that GnRH-(1-5) retards the cellular migration of neural cells via
GPR173 (71–73). By contrast, GnRH-(1-5) may stimulate cellular
migration and invasion of the extracellular matrix in endometrial
cells via GPR101 (70,77).
These studies support the idea that GnRH-(1-5) represents
another layer of regulatory complexity in tissues where GnRH is
also produced. The identification of an endogenous ligand to an
orphan GPCR is important because these receptors may have thera-
peutic potential (74). Furthermore, the identification of a GPCR that
binds GnRH-(1-5) may help resolve some of the current quandaries
regarding the actions of GnRH (agonist/antagonist) and enhance
our understanding in the evolution of peptide metabolism and
processing.
Role of GnIH in avian reproductive system; regulation of
GnIH by photoperiod and stress and the effects of these
changes on reproductive behaviours
Although GnRH and its metabolite, GnRH-(1-5), are known for pro-
moting reproduction-related functions in the HPG axis, a more
recently discovered hormone has been implicated as a potential
brake on the HPG system. GnIH has received attention because of
its role in the inhibition of activity of components of the HPG axis,
including a reduction of sexual behaviour (78–86). Despite a great
deal of investigation into its specific functions and the factors that
Kiss Neuron
E2ERα
3β-HSD
ERα
E2
Kisspeptin
mRNA/protein
MAPK p
p
Src
PR
+
+
+
Kisspeptin
release
GPR54
GnR
H
E2
ERα
neuroP
PREG
O
[Ca2+]i
LH
P450scc
Gonadotroph
Astrocyte
mGluR1a
Fig. 1. A model showing proposed actions of oestradiol (E
2
) on hypothalamic cells. In kisspeptin (Kiss1) neurones, E
2
acts at both membrane and nuclear
oestrogen receptors. During di-oestrus, classical nuclear E
2
signalling induces progesterone receptor (PR) expression in Kiss1 neurones in the rostral periventric-
ular nucleus of the third ventricle (RP3V). On pro-oestrus, rising E
2
leads to transactivation of mGluR1a in astrocytes, which increases [Ca
2+
]
i
, leading to the
conversion of cholesterol to pregnenolone (PREG) by the P450scc enzyme and the conversion of PREG to progesterone (neuroP) by 3b-hydroxysteroid dehydro-
genase (HSD). Simultaneously, E
2
activates an oestrogen receptor (ER)a-mGluR1a complex in neurones leading to the expression of Kiss1. Newly synthesised
neuroP diffuses out of the astrocytes and activates E
2
-induced PR, which has been trafficked to the Kiss1 neuronal membrane. This leads to a series of events
culminating in Kiss1 secretion onto GPR54 expressing gonadotrophin-releasing hormone (GnRH) neurones. Signalling through PR in Kiss1 neurones induces
Kiss1 release, activating GnRH neurones and triggering the E
2
-induced luteinising hormone (LH) surge from anterior pituitary gonadotrophs. MAPK, mitogen-
activated protein kinase.
©2016 British Society for NeuroendocrinologyJournal of Neuroendocrinology, 2016, 28, 10.1111/jne.12405
Peripheral and central hormonal control of reproduction 3of12
regulate GnIH, the full range of actions of GnIH within the central
nervous system remain unknown. At present, we know that, in
birds, GnIH projects to GnIH receptor-expressing GnRH-I and -II
neurones in addition to the median eminence (84,87). In several
species of mammals, GnIH projects to and also influences the activ-
ity of GnRH neurones (85,88), as well as the external layer of the
median eminence (88–92), although this latter finding remains dis-
puted (85,93). There are GnIH projections to multiple other brain
areas (e.g. brainstem) and possibly to the spinal cord (84,93),
although the function of GnIH in these extra-hypothalamic areas
remains obscure. The GnIH content of the brain is influenced by
changes in day length and the associated changing melatonin sig-
nal in seasonal breeders (84,94–99). In birds, despite the influence
of GnIH on GnRH neurones, it appears that GnIH does not influ-
ence the termination of reproduction at the end of the breeding
season. Rather, it is more likely that GnIH plays a role in temporary
reproductive suppression within the breeding season in response to
different physiological stimuli, such as stress (84,100–102).
The action of GnIH is not restricted to the brain and the anterior
pituitary gland. GnIH and its receptor (GPR147) are synthesised
in the gonads of both sexes of all vertebrates studied to date
(103–108). Furthermore, in birds, GnIH-producing neurones in the
brain project to the pars nervosa, suggesting that GnIH is released
directly into the bloodstream (G. Bentley, unpublished observations).
If confirmed, then not only can locally produced GnIH act within
the gonads, but also neurally produced GnIH could be released to
the general circulation and act upon peripheral targets.
It is possible that GnIH-producing neurones can be subdivided
into heterogenous subpopulations that respond to unique environ-
mental and physiological cues. For example, GnIH neurones express
melatonin receptor (MelR) and glucocorticoid receptor (GR) mRNA.
However, not all of the GnIH neurones express MelR or GR (98,109)
and it is not known whether single GnIH neurones can express
both MelR and GR, suggesting that there could be MelR- and GR-
specific subpopulations of GnIH neurones, each with potentially dis-
tinct functions. Thus, it remains to be determined whether or how
melatonin and glucocorticoids interact to influence GnIH action
within the brain.
In birds and mammals, melatonin and corticosterone can act on
the gonadal GnIH system. This suggests the possibility that the
neural and gonadal GnIH systems could differentially respond to
hormones and, together, could coordinate a response to circulating
hormones (perhaps via direct innervation of the gonad). Unfortu-
nately, only in vitro preparations can be used to answer this ques-
tion. Without separating the gonads from the blood circulation and
from potential neural input, it is impossible to determine gonadal
responses to a changing hormonal environment, especially if GnIH
is present in circulating blood. However, in vivo studies in this area
pGlu
His
Trp
Ser
Tyr Gly
Leu
Arg
Pro
Gly
GnRH
pGlu
His
Ser
Tyr
EP24.15
NH2
GnRH-(1-5)
GPR101
COOH
GPR173
NH2
COOH
Trp
Fig. 2. Gonadotropin-Releasing Hormone (GnRH) peptide processing and action. The decapeptide, GnRH, is processed extracellularly to form the metabolite,
GnRH-(1-5) by the zinc metalloendopeptidase, EC3.4.24.15 (EP24.15; 66, 73). The metabolite, GnRH-(1-5), exerts is biological activities via 2 putative receptors,
the G-protein coupled receptors (GPR) GPR101 and GPR173 (70, 71).
©2016 British Society for Neuroendocrinology Journal of Neuroendocrinology, 2016, 28, 10.1111/jne.12405
4of12 L. M. Rudolph et al.
could also be very informative, especially if localised blockade of
GnIH receptor could be induced in the gonads.
GnIH responses to chronic stress have been documented in male
and female rats, with a significant impact on reproduction
(109–111). To date, there has been only one study on chronic stress
effects upon GnIH in birds with sex-specific effects of treatment.
Female European starlings (Sturnus vulgaris) exhibited increased
ovarian GnIH expression compared to their nonstressed counter-
parts and were also reported not to ovulate, whereas nonstressed
animals did (111). Acute stressors can certainly influence the avian
GnIH system, although these effects appear to depend on the spe-
cies, the time of year, the sex of the bird and the stressor
(112–114). In addition, some stressors influence the gonads directly
rather than via neural GnIH (112). The same is true for chronic
housing stress in European starlings, as noted above (111). Thus, it
is clear that neural and GnIH systems can respond differently to
any particular stressor, regardless of whether it is acute or chronic.
Further studies in this area should determine the response of gona-
dal and neural GnIH systems to stressors and hormones, and
should also assess communication between these GnIH systems in
a variety of species.
Local regulation of gonadal function
Autocrine and paracrine regulation of testicular function:
molecular pathways involved in testis pathophysiology
leading to infertility
Gonadotrophins are key regulators of male gonadal function. LH
and follicle-stimulating hormone (FSH) released from the pituitary
reach the testis and exert their effects through receptors located in
the plasma membrane of Leydig and Sertoli cells, respectively
(115,116). In addition, local factors and hormones influence testicu-
lar function via paracrine and autocrine mechanisms. Several mole-
cules that reach the testis and/or are locally produced in the gonad
regulate the activity of different cell types (e.g. Leydig cells, Sertoli
cells, mast cells, macrophages, myofibroblasts), include peptides
(117), neurotransmitters (118), neurohormones (119), cytokines
(120) and prostaglandins (PGs) (121).
In this context, the neurohormones serotonin (122), melatonin
(123) and corticotrophin-releasing hormone (CRH) (124) act in the
testes as important negative regulators of cAMP and androgen pro-
duction. Serotonin, melatonin and CRH can be produced within the
CNS and secreted into peripheral circulation, or locally synthesised
in the testes (125,126). Melatonin and also serotonin inhibit
steroidogenesis via their 5-HT
2
receptor- and Mel1a receptor-
mediated signalling pathways, which influence CRH centrally
(125,127) and in the testes (127,128). This CRH-mediated inhibition
of steroid production occurs through the activation of tyrosine
phosphatases, which reduces the phosphorylation of extracellular
regulated kinase (ERK) and c-Jun N-terminal kinase, and subse-
quently down-regulates c-jun, c-fos and steroid acute regulatory
protein (StAR), thereby inhibiting testosterone production (128).
Melatonin has been postulated to have a physiological role as a
paracrine signalling molecule, directly regulating the production of
factors (e.g. immune, interleukin-2) in its immediate vicinity (129).
Recent observations show that melatonin modulates local cellular
activity in testicular immune cells, inducing the expression of
antioxidant enzymes and reducing the generation of reactive oxy-
gen species in mast cells. In testicular macrophages, melatonin inhi-
bits cell proliferation, the expression of proinflammatory cytokines,
interleukin-1band tumour necrosis factor a, and PG production
(130). PGs are derived from arachidonic acid by the action of indu-
cible isoenzyme cyclooxygenase (COX). In testicular biopsies of men
with impaired spermatogenesis, COX-2 is expressed in immune cells,
highlighting their relevance in testicular inflammation associated
with idiopathic infertility (131). Furthermore, Leydig and Sertoli cells
also produce PGs and express several prostanoid receptors
(132,133), suggesting autocrine/paracrine action in testicular
somatic cells.
PGD2 has a stimulatory effect on basal testosterone production
in Leydig cells (134), whereas PGF2aexerts an inhibitory role in the
expression of the StAR and 17b-hydroxysteroid dehydrogenase
(HSD), as well as in the synthesis of testosterone induced by human
chorionic gonadotrophin (hCG)/LH (133), demonstrating that the
role of PGs on steroidogenesis, spermatogenesis and ultimately fer-
tility depends on the specific PG in question.
Recent research indicates that multiple local signals influence
testicular physiology and are involved in the pathogenesis or main-
tenance of human infertility. Notably, male infertility results from
endocrine dysfunctions associated with the hypothalamic-pituitary-
testicular axis only in a small number of cases (135), suggesting
the source of infertility likely occurs within local, intra-testicular
pathways. Thus, new insights about how cell–cell interactions
within the testes affect testicular function and fertility will con-
tribute to the understanding of male reproductive physiopathology,
and future studies focusing on testicular paracrine and autocrine
interactions may lead to new therapeutic approaches to idiopathic
male infertility.
Follicular development, corpus luteum and progesterone
regulation of ovarian vascularisation and molecular
pathways involved
Similar to testicular functions including spermatogenesis and
steroidogenesis, ovarian follicular development and regression is a
continuous and cyclic process that depends on a number of endo-
crine, paracrine and autocrine signals. In healthy tissues, physiologi-
cal angiogenesis is mainly limited to the reproductive system. The
ovarian vasculature is closely associated with preovulatory follicle
and corpus luteum during the ovarian cycle and is one of the few
sites where nonpathological development and regression of blood
vessels occurs in the adult. Recently, local factors such as vascular
endothelial growth factor A (VEGF-A) and angiopoietins, which act
specifically on vascular endothelial cells or pericytes and smooth
muscle to control angiogenesis or angiolysis, were identified in the
growing follicle and corpus luteum of several species, including
humans (136).
VEGF-A is a key angiogenic factor involved in the formation of
new blood vessels within many tissues. It is required to initiate the
©2016 British Society for NeuroendocrinologyJournal of Neuroendocrinology, 2016, 28, 10.1111/jne.12405
Peripheral and central hormonal control of reproduction 5of12
formation of new immature vessels by promoting endothelial cell
proliferation and vascular permeability. Inhibition of VEGF-A and
angiopoietin 1 (ANGPT1) action in rat ovaries by intrabursal admin-
istration of VEGF-A-Trap or ANGPT1 antibodies, respectively, pro-
duces an imbalance in the ratio of anti-apoptotic : pro-apoptotic
proteins leading to greater follicular atresia (137,138). In addition,
VEGF-A prevents apoptosis and stimulates the proliferation of gran-
ulosa and theca cells of antral follicles through a direct interaction
with its KDR receptor localised in granulosa cells, a pathway that
involves phosphoinositide 3-kinase (PI3K)/AKT (139). Furthermore,
in vitro studies performed in early antral follicles and granulosa cell
cultures isolated from rat demonstrate that VEGF acts directly on
follicular cells synergistically with FSH and E
2
, preventing apoptosis
and stimulating proliferation, thus promoting follicular development
and the selection of the follicle to ovulate (140). Such work
reported a direct role for VEGF in early antral follicles mediated by
the PI3K/AKT and ERK1/2 pathways, besides the classical and well
known proangiogenic function. Together, these data support the
notion that angiogenic factors have an important role in controlling
ovarian function.
In vitro studies have shown that Notch signalling is critical for
the survival of luteal cells isolated from pregnant rats (141). Local
Notch inhibition decreases progesterone levels and cell survival,
confirming that Notch has a direct action on both steroidogenesis
and luteal viability (141). The Notch signalling pathway is a cell–cell
communication pathway that is evolutionarily conserved from Dro-
sophila to humans. To date, four different Notch receptors (Notch1,
2, 3 and 4) and five different ligands (Jagged-1 and -2 and DLL-1 -
3 and -4) have been identified in mammals. This Notch system reg-
ulates cell fate, proliferation and death. The Notch genes encode
transmembrane receptors, which, upon binding their ligand, are
cleaved, releasing the intracellular domain. The intracellular portion
of the receptor translocates to the nucleus to act as a transcrip-
tional coactivator, regulating cell fate genes (142).
Moreover, in the rat, there is an interaction between the Notch
signalling pathway and progesterone that maintains the functional-
ity of the corpus luteum (143). Notch signalling augments P450scc
synthesis, leading to an increased synthesis of progesterone, which
in turn regulates the activated intracellular Notch domain. Thus,
Notch induces progesterone production in vitro through the activa-
tion of cytochrome P450 cholesterol side chain cleavage enzyme
(P450scc) and decreases apoptosis-mediated cell death. This is the
first evidence that there is cross-talk between the Notch signalling
system and progesterone, which increases the survival of luteal
cells. Also, the Notch/PI3K/AKT signalling pathway might be inter-
acting with progesterone, intensifying the survival role of this hor-
mone in luteal cells. Nevertheless, future studies are required to
thoroughly investigate this newly discovered Notch-progesterone
relationship and how it contributes to ovarian function and repro-
duction as a whole.
Ovarian kisspeptin and its role in follicular development
Reproduction in females requires an LH surge, which is centrally
regulated by KP. However, KP is found in many peripheral organs
(144,145), in particular, the ovary, which expresses KP and its
receptor, GPR54, suggesting a role for KP in the peripheral control
of reproductive events. KP expression in the ovary fluctuates
throughout the oestrous cycle, strongly suggesting that it may be
involved locally in the ovulatory cycle and luteinisation (146–148);
but see also (28). However, the mechanisms of action of KP in the
ovary, such as paracrine or autocrine functions remain largely
unknown.
A recent study demonstrated that intraovarian administration of
a KP antagonist (p234) delays vaginal opening and alters the oes-
trous cycle in rats (147). Additionally, local administration of exoge-
nous KP decreases antral follicle and corpora lutea number in
fertile and subfertile rats, which was reversed by p234 treatment,
suggesting that KP also participates in both follicular development
and ovulation at the level of the ovary (149). Moreover, during ovu-
lation in humans and nonhuman primates, ovarian KP and GPR54
mRNA increases with other ovulation-associated genes, such as
COX-2 and progesterone receptor. The ovarian administration of the
COX-2 inhibitor, indomethacin, disrupted the ovulatory process in
rats, supporting the idea of a local role of KP and GRP54 in ovula-
tion (150). It appears that KP regulates progesterone secretion from
luteal cells as well. In isolated chicken granulosa cells, KP stimulates
progesterone secretion, possibly by directly altering levels of
steroidogenic enzymes, including StAR, P450scc, which converts
cholesterol to pregnenolone, and 3b-HSD (151), which converts
pregnenolone to progesterone. Similarly, in rat luteal cells, KP
increased progesterone production via ERK1/2 signalling and
increased the expression of StAR and CYP11A mRNA (152). Further-
more, administration of a GPR54 antagonist, p234, inhibited pro-
gesterone secretion in granulosa cell cultures treated with hCG,
implicating KP in the luteinisation of granulosa cells (148). Together,
these data suggest a potential role of KP in the local control of
ovarian function, potentially via progesterone synthesis. These and
future studies involving paracrine and autocrine actions of ovarian
KP will clarify the molecular mechanisms involved in the regulation
of follicular development and ovulation during reproductive life and
ovarian ageing.
Although a decreased follicular pool indicates physiological age-
ing of the ovary (153), an increased rate of follicular loss is also a
pathology that affects the follicular reserve pool, and thereby fertil-
ity, in humans and other mammals (154). Reproductive ageing in
women begins with shortened menstrual cycles, smaller increases
in FSH and decreased levels of inhibin (155), which results in accel-
erated follicular growth and premature exhaustion of the follicular
pool. One of the mechanisms involved in ovarian ageing is
increased sympathetic nerve activity. Ovaries of postmenopausal
women (51 years old) have a higher density of innervation com-
pared to age-matched controls (156,157). In the rat, reproductive
ageing is associated with increased ovarian sympathetic activity,
which is strongly correlated with the spontaneous appearance of
follicular cysts and a loss of preantral follicles (158,159). Indeed,
the highest sympathetic innervation is found in postmenopausal
women, suggesting a correlation between ageing-induced infertility
and sympathetic nerve activation. Recent findings indicate that
sympathetic innervation may be controlling age-induced infertility
©2016 British Society for Neuroendocrinology Journal of Neuroendocrinology, 2016, 28, 10.1111/jne.12405
6of12 L. M. Rudolph et al.
via regulation of KP because ovarian sympathectomy diminishes KP
levels (A. Paredes, unpublished observations). Additionally, during
reproductive ageing, KP expression in the ovary increases from the
subfertile to infertile period and is directly correlated with the
increase in ovarian norepinephrine observed with ageing (149,158),
suggesting that KP may be directly controlled by sympathetic inner-
vation of the ovary (147), as well as supporting the idea that KP is
regulated by the adrenergic system and that both the adrenergic
system and KP participate in the local regulation of follicle develop-
ment and ovulation during reproductive ageing. Furthermore, KP is
involved in follicular dynamics: intraovarian administration of KP
produced an increase in the numbers of corpora lutea and type III
follicles in fertile and subfertile periods, which was reversed by KP
receptor antagonism. Future studies should address the potential
autocrine and paracrine roles of KP in the ovary, specifically the
interaction of KP, steroidogenic pathways and sympathetic innerva-
tion and how they relate to reproductive outcomes across the
lifespan.
Morphological changes in ARH initiated by oestradiol
membrane signalling that mediate lordosis behaviour
Another key component to reproduction in rodents is female sexual
receptivity, which is mediated by E
2
-dependent alterations in
hypothalamic neuronal structure. Although the molecular bases of
E
2
-dependent facilitation of female sexual receptivity have more
recently been described in detail, the understanding that steroid
hormones exert behavioural effects via changes in neural morphol-
ogy is a well established phenomenon. The most well known exam-
ple of E
2
-induced changes in dendritic structure regulating
memory-related behaviour is from the hippocampus (160), whereas
E
2
-induced changes in dendrites in the hypothalamus have also
been known for some time (161). Indeed, changes in dendritic mor-
phology are critical for the lordosis-regulating circuit (162), which
extends from the ARH to the medial preoptic nucleus (MPN), to the
ventromedial nucleus of the hypothalamus (VMH). Recent studies
have begun to clarify the molecular mechanisms by which morpho-
logical changes in the ARH-MPN-VMH circuit allow for expression
of lordosis behaviour. The primary step of E
2
signalling in the ARH
occurs via ERatransactivation of mGluR1a, which initiates morpho-
logical changes that are coincident with and required for the dis-
play of lordosis behaviour. Within 4 h after E
2
treatment, immature,
filapodia-like dendritic spines are formed in the ARH (162). Twenty-
four hours after E
2
treatment, there is a shift in the proportion of
dendritic spines, with a decrease in filapodia and a concomitant
increase in mature, mushroom-shaped spines (162). The formation
of new spines is necessary for the E
2
-induced lordosis because
blocking spine formation significantly reduces the expression of
sexual receptivity (162).
Although it appears that spinogenesis is initiated by the action
of E
2
at membrane ERa, it is unclear what molecular mechanisms
underlie spine maturation. Evidence from other circuits suggests a
role for the G-protein coupled ER, GPR30, in spine maturation and
stabilisation. GPR30 is localised in spine heads, associates with
PSD-95, and is regulated by E
2
(163,164). In the dorsal
hippocampus, the GPR30 agonist, G1, increases PSD-95 immunore-
activity, suggesting a role for GPR30 in spine maturation (164).
Indeed, this receptor has been implicated in the initiation of lordo-
sis behaviour on the basis that the partial GPR30 agonist but ERa
antagonist, ICI 182,780, facilitates lordosis in E
2
-primed nonrecep-
tive rats (165). Other studies suggest there could be a role for the
STX-activated G
q
-coupled membrane ER in the ARH-MPN circuit
mediating sexual receptivity. STX is a tamoxifen analogue that does
not bind to classical ERaor GPR30 but is blocked by the ER antag-
onist ICI 182,780 and has pharmacological profile similar to those
of the ERa-specific agonist, PPT (166–168). STX treatment induces
l-opioid receptor (MOR) internalisation in the MPN and facilitates
lordosis behaviour (169). Alternatively, spine maturation could be
mediated by extra-neuronal mechanisms, such as astrocytic contact
with neurones, which alters dendritic spine formation and stabilisa-
tion (170).
Additionally, it is unclear whether E
2
induces spinogenesis in the
same population of neurones in the ARH that express ERa, the
neuropeptide Y (NPY) neurones, which are the initial site of the
action of E
2
in the ARH-MPN-VMH circuit, or whether E
2
is acting
transsynaptically to induce spines on pro-opiomelanocortin (POMC)
neurones, which release b-endorphin onto MORs in the MPN.
Recent data suggest that the NPY neurones and not POMC neu-
rones undergo spinogenesis, suggesting that spine formation occurs
directly within the neurones where initial E
2
activation of ERa
occurs (171). Regardless of the site of spinogenesis within the ARH,
it is clear that spine maturation in this nucleus is coincident with
lordosis behaviour, and also that blocking spinogenesis here reduces
female sexual receptivity. To a first approximation, the timeline
from E
2
treatment to the presence of mature dendritic spines is
known. However, the time when fully functional synapses appear
remains to be determined. Within 1 h of E
2
treatment, cofilin is
deactivated via phosphorylation, which permits spinogenesis (162),
and, in the MPN, MOR is activated/internalised, indicating that the
ARH to MPN part of the circuit is functional (172). At 4 h post-E
2
treatment, filapodial spines are present, although these thin, labile
spines are not considered to mediate functional synapses (173). At
20 h after E
2
treatment, the first time point when lordosis beha-
viour can be elicited with supplemental hormone treatment, there
is an increase in the proportion of mushroom spines that are gen-
erally assumed to be indicative of functional synapses (162) and
that contain the machinery required for synaptic transmission (e.g.
PSD-95). Future studies should address the time course of this E
2
-
dependent spine maturation and the potential involvement of non-
traditional ER in this process.
Conclusions
Taken together, these recent findings highlight both the redundancy
and complexity of the hormonal control of reproduction: what was
once considered to be a simple, direct circuit with a handful of
steroid hormones and cognate receptors is continually updated with
novel hormone regulators and mechanisms of hormone synthesis
and action. However, the classical aspects of gonadal hormone con-
trol of reproduction remain intact, demonstrating that there are
©2016 British Society for NeuroendocrinologyJournal of Neuroendocrinology, 2016, 28, 10.1111/jne.12405
Peripheral and central hormonal control of reproduction 7of12
multiple levels of control of the HPG axis, both centrally and
peripherally. Future studies will likely only add to this increasingly
complex circuit that regulates reproduction.
Disclaimer
The opinions or assertions contained herein are the private ones of
the authors and are not to be construed as official or reflecting the
views of the Department of Defense or the Uniformed Services
University of the Health Sciences.
Received 14 January 2016,
revised 25 May 2016,
accepted 20 June 2016
References
1 Kuo J, Micevych P. Neurosteroids, trigger of the LH surge. J Steroid
Biochem Mol Biol 2012; 131: 57–65.
2 Brom GM, Schwartz NB. Acute changes in the estrous cycle following
ovariectomy in the golden hamster. Neuroendocrinology 1968; 3: 366–
377.
3 Ferin M, Tempone A, Zimmering PE, Van de Wiele RL. Effect of anti-
bodies to 17beta-estradiol and progesterone on the estrous cycle of
the rat. Endocrinology 1969; 85: 1070–1078.
4 Labhsetwar AP. Role of estrogens in ovulation: a study using the
estrogen-antagonist, I.C.I. 46,474. Endocrinology 1970; 87: 542–551.
5 Micevych P, Sinchak K, Mills RH, Tao L, LaPolt P, Lu JK. The luteinizing
hormone surge is preceded by an estrogen-induced increase of
hypothalamic progesterone in ovariectomized and adrenalectomized
rats. Neuroendocrinology 2003; 78: 29–35.
6 Hibbert ML, Stouffer RL, Wolf DP, Zelinski-Wooten MB. Midcycle
administration of a progesterone synthesis inhibitor prevents ovulation
in primates. Proc Natl Acad Sci USA 1996; 93: 1897–1901.
7 Remohi J, Balmaceda JP, Rojas FJ, Asch RH. The role of pre-ovulatory
progesterone in the midcycle gonadotrophin surge, ovulation and sub-
sequent luteal phase: studies with RU486 in rhesus monkeys. Hum
Reprod 1988; 3: 431–435.
8 DePaolo LV. Attenuation of preovulatory gonadotrophin surges by
epostane: a new inhibitor of 3 beta-hydroxysteroid dehydrogenase. J
Endocrinol 1988; 118: 59–68.
9 Mahesh VB, Brann DW. Interaction between ovarian and adrenal ster-
oids in the regulation of gonadotropin secretion. J Steroid Biochem
Mol Biol 1992; 41: 495–513.
10 DePaolo LV, Barraclough CA. Dose dependent effects of progesterone on
the facilitation and inhibition of spontaneous gonadotropin surges in
estrogen treated ovariectomized rats. Biol Reprod 1979; 21: 1015–1023.
11 Petersen SL, Keller ML, Carder SA, McCrone S. Differential effects of
estrogen and progesterone on levels of POMC messenger RNA levels in
the arcuate nucleus relationship to the timing of LH surge release. J
Neuroendocrinol 1993; 5: 643–648.
12 Petersen SL, McCrone S, Keller M, Shores S. Effects of estrogen and
progesterone on luteinizing hormone-releasing hormone messenger
ribonucleic acid levels: consideration of temporal and neuroanatomical
variables. Endocrinology 1995; 136: 3604–3610.
13 Chappell PE, Levine JE. Stimulation of gonadotropin-releasing hormone
surges by estrogen. I. Role of hypothalamic progesterone receptors.
Endocrinology 2000; 141: 1477–1485.
14 Micevych PE, Sinchak K. The neurosteroid progesterone underlies estro-
gen positive feedback of the LH surge. Front Endocrinol 2011; 2: 90.
15 Stephens SB, Tolson KP, Rouse MLJ, Poling MC, Hashimoto-Parktya M,
Mellon PL, Kauffman AS. Absent progesterone signaling in kisspeptin
neurons disrupts the LH surge and impairs fertility in female mice.
Endocrinology 2015; 156: 3091–3097.
16 Herbison AE, Theodosis DT. Localization of oestrogen receptors in pre-
optic neurons containing neurotensin but not tyrosine hydroxylase,
cholecystokinin or luteinizing hormone-releasing hormone in the male
and female rat. Neuroscience 1992; 50: 283–298.
17 Shivers BD, Harlan RE, Morrell JI, Pfaff DW. Absence of oestradiol con-
centration in cell nuclei of LHRH-immunoreactive neurones. Nature
1983; 304: 345–347.
18 Wintermantel TM, Campbell RE, Porteous R, Bock D, Grone HJ, Todman
MG, Korach KS, Greiner E, Perez CA, Schutz G, Herbison AE. Definition
of estrogen receptor pathway critical for estrogen positive feedback to
gonadotropin-releasing hormone neurons and fertility. Neuron 2006;
52: 271–280.
19 Han SK, Gottsch ML, Lee KJ, Popa SM, Smith JT, Jakawich SK, Clifton
DK, Steiner RA, Herbison AE. Activation of gonadotropin-releasing hor-
mone neurons by kisspeptin as a neuroendocrine switch for the onset
of puberty. J Neurosci 2005; 25: 11349–11356.
20 Smith JT, Cunningham MJ, Rissman EF, Clifton DK, Steiner RA. Regula-
tion of Kiss1 gene expression in the brain of the female mouse.
Endocrinology 2005; 146: 3686–3692.
21 Liu X, Lee K, Herbison AE. Kisspeptin excites gonadotropin-releasing hor-
mone neurons through a phospholipase C/calcium-dependent pathway
regulating multiple ion channels. Endocrinology 2008; 149: 4605–4614.
22 Dumalska I, Wu M, Morozova E, Liu R, van den Pol A, Alreja M. Excita-
tory effects of the puberty-initiating peptide kisspeptin and group I
metabotropic glutamate receptor agonists differentiate two distinct
subpopulations of gonadotropin-releasing hormone neurons. J Neurosci
2008; 28: 8003–8013.
23 Zhang C, Roepke TA, Kelly MJ, Ronnekleiv OK. Kisspeptin depolarizes
gonadotropin-releasing hormone neurons through activation of TRPC-
like cationic channels. J Neurosci 2008; 28: 4423–4434.
24 Pielecka-Fortuna J, Chu Z, Moenter SM. Kisspeptin acts directly and
indirectly to increase gonadotropin-releasing hormone neuron activity
and its effects are modulated by estradiol. Endocrinology 2008; 149:
1979–1986.
25 Clarkson J, D’Anglemont de Tassigny X, Moreno AS, Colledge WH, Her-
bison AE. Kisspeptin-GPR54 signaling is essential for preovulatory
gonadotropin-releasing hormone neuron activation and the luteinizing
hormone surge. J Neurosci 2008; 28: 8691–8697.
26 Navarro VM, Castellano JM, Fernandez-Fernandez R, Barreiro ML, Roa
J, Sanchez-Criado JE, Aguilar E, Dieguez C, Pinilla L, Tena-Sempere M.
Developmental and hormonally regulated messenger ribonucleic acid
expression of KiSS-1 and its putative receptor GPR54 in rat hypothala-
mus and potent LH releasing activity of KiSS-1 peptide. Endocrinology
2004; 145: 4565–4574.
27 Seminara SB. Metastin and its G protein-coupled receptor, GPR54: crit-
ical pathway modulating GnRH secretion. Front Neuroendocrinol 2005;
26: 131–138.
28 Kirilov M, Clarkson J, Liu X, Roa J, Campos P, Porteous R, Schutz G,
Herbison AE. Dependence of fertility on kisspeptin-Gpr54 signaling at
the GnRH neuron. Nat Commun 2013; 4: 2492.
29 Kinoshita M, Tsukamura H, Adachi S, Matsui H, Uenoyama Y, Iwata K,
Yamada S, Inoue K, Ohtaki T, Matsumoto H, Maeda K. Involvement of
central metastin in the regulation of preovulatory luteinizing hormone
surge and estrous cyclicity in female rats. Endocrinology 2005; 146:
4431–4436.
30 Smith JT, Popa SM, Clifton DK, Hoffman GE, Steiner RA. Kiss1 neurons
in the forebrain as central processors for generating the preovulatory
luteinizing hormone surge. J Neurosci 2006; 26: 6687–6694.
©2016 British Society for Neuroendocrinology Journal of Neuroendocrinology, 2016, 28, 10.1111/jne.12405
8of12 L. M. Rudolph et al.
31 Chappell PE, Lee J, Levine JE. Stimulation of gonadotropin-releasing
hormone surges by estrogen. II. Role of cyclic adenosine 3’5’-mono-
phosphate. Endocrinology 2000; 141: 1486–1492.
32 Chappell PE, Schneider JS, Kim P, Xu M, Lydon JP, O’Malley BW, Levine
JE. Absence of gonadotropin surges and gonadotropin-releasing hor-
mone self- priming in ovariectomized (OVX), estrogen (E2)-treated, pro-
gesterone receptor knockout (PRKO) mice. Endocrinology 1999; 140:
3653–3658.
33 Zhang J, Yang L, Lin N, Pan X, Zhu Y, Chen X. Aging-related changes
in RP3V kisspeptin neurons predate the reduced activation of GnRH
neurons during the early reproductive decline in female mice. Neuro-
biol Aging 2014; 35: 655–668.
34 Smith JT, Clay CM, Caraty A, Clarke IJ. KiSS-1 messenger ribonucleic
acid expression in the hypothalamus of the ewe is regulated by sex
steroids and season. Endocrinology 2007; 148: 1150–1157.
35 Soma KK, Sinchak K, Lakhter A, Schlinger BA, Micevych PE. Neurosteroids
and female reproduction: estrogen increases 3beta-HSD mRNA and
activity in rat hypothalamus. Endocrinology 2005; 146: 4386–4390.
36 Mittelman-Smith MA, Wong AM, Kathiresan AS, Micevych PE. Classical
and membrane-initiated estrogen signaling in an in vitro model of
anterior hypothalamic kisspeptin neurons. Endocrinology 2015; 156:
2162–2173.
37 Zuloaga DG, Yahn SL, Pang Y, Quihuis AM, Oyola MG, Reyna A, Thomas
P, Handa RJ, Mani SK. Distribution and estrogen regulation of mem-
brane progesterone receptor-beta in the female rat brain. Endocrinol-
ogy 2012; 153: 4432–4443.
38 Thomas P, Pang Y, Dong J, Groenen P, Kelder J, de Vlieg J, Zhu Y,
Tubbs C. Steroid and G protein binding characteristics of the seatrout
and human progestin membrane receptor alpha subtypes and their
evolutionary origins. Endocrinology 2007; 148: 705–718.
39 Tang YT, Hu T, Arterburn M, Boyle B, Bright JM, Emtage PC, Funk WD.
PAQR proteins: a novel membrane receptor family defined by an
ancient 7-transmembrane pass motif. J Mol Evol 2005; 61: 372–380.
40 Moussatche P, Lyons TJ. Non-genomic progesterone signalling and its
non-canonical receptor. Biochem Soc Trans 2012; 40: 200–204.
41 Thomas P, Pang Y. Membrane progesterone receptors: evidence for
neuroprotective, neurosteroid signaling and neuroendocrine functions
in neuronal cells. Neuroendocrinology 2012; 96: 162–171.
42 Pang Y, Dong J, Thomas P. Characterization, neurosteroid binding and
brain distribution of human membrane progesterone receptors delta and
epsilon (mPRdelta and mPR{epsilon}) and mPRdelta involvement in neu-
rosteroid inhibition of apoptosis. Endocrinology 2013; 154: 283–295.
43 Yoshikuni M, Nagahama Y. Involvement of an inhibitory G-protein in
the signal transduction pathway of maturation-inducing hormone (17
alpha,20 beta-dihydroxy-4-pregnen-3-one) action in rainbow trout
(Oncorhynchus mykiss) oocytes. Dev Biol 1994; 166: 615–622.
44 Karteris E, Zervou S, Pang Y, Dong J, Hillhouse EW, Randeva HS, Tho-
mas P. Progesterone signaling in human myometrium through two
novel membrane G protein-coupled receptors: potential role in func-
tional progesterone withdrawal at term. Mol Endocrinol 2006; 20:
1519–1534.
45 Bashour NM, Wray S. Progesterone directly and rapidly inhibits GnRH
neuronal activity via progesterone receptor membrane component 1.
Endocrinology 2012; 153: 4457–4469.
46 Hanna R, Pang Y, Thomas P, Zhu Y. Cell-surface expression, progestin
binding, and rapid nongenomic signaling of zebrafish membrane pro-
gestin receptors alpha and beta in transfected cells. J Endocrinol 2006;
190: 247–260.
47 Ashley RL, Clay CM, Farmerie TA, Niswender GD, Nett TM. Cloning and
characterization of an ovine intracellular seven transmembrane recep-
tor for progesterone that mediates calcium mobilization. Endocrinology
2006; 147: 4151–4159.
48 Krietsch T, Fernandes MS, Kero J, Losel R, Heyens M, Lam EW, Huh-
taniemi I, Brosens JJ, Gellersen B. Human homologs of the putative G
protein-coupled membrane progestin receptors (mPRalpha, beta, and
gamma) localize to the endoplasmic reticulum and are not activated
by progesterone. Mol Endocrinol 2006; 20: 3146–3164.
49 Paaske LK, Chuon T, Micevych P, Sinchak K. AVPV Kisspeptin Neurons
Mediate Neuroprogesterone Induction of the Leuteinizing Hormone
Surge. Washington, DC: Society for Neuroscience, 2014.
50 Gorbman A, Sower SA. Evolution of the role of GnRH in animal (Meta-
zoan) biology. Gen Comp Endocrinol 2003; 134: 207–213.
51 Ojeda SR, Lomniczi A. Puberty in 2013: unraveling the mystery of pub-
erty. Nat Rev Endocrinol 2014; 10: 67–69.
52 Semaan SJ, Tolson KP, Kauffman AS. The development of kisspeptin cir-
cuits in the Mammalian brain. Adv Exp Med Biol 2013; 784: 221–252.
53 Yeo SH. Neuronal circuits in the hypothalamus controlling gonadotro-
phin-releasing hormone release: the neuroanatomical projections of
kisspeptin neurons. Exp Physiol 2013; 98: 1544–1549.
54 Kelly MJ, Zhang C, Qiu J, Ronnekleiv OK. Pacemaking kisspeptin neu-
rons. Exp Physiol 2013; 98: 1535–1543.
55 Okamura H, Tsukamura H, Ohkura S, Uenoyama Y, Wakabayashi Y,
Maeda K. Kisspeptin and GnRH pulse generation. Adv Exp Med Biol
2013; 784: 297–323.
56 Smith JT. Sex steroid regulation of kisspeptin circuits. Adv Exp Med Biol
2013; 784: 275–295.
57 Terasawa E, Guerriero KA, Plant TM. Kisspeptin and puberty in mam-
mals. Adv Exp Med Biol 2013; 784: 253–273.
58 DePaolo LV, King RA, Carillo AJ. In vivo and in vitro examination of an
autoregulatory mechanism for luteinizing hormone-releasing hormone.
Endocrinology 1987; 120: 272–279.
59 Sarkar DK. In vivo secretion of LHRH in ovariectomized rats is regu-
lated by a possible autofeedback mechanism. Neuroendocrinology
1987; 45: 510–513.
60 Sarkar DK, Chiappa SA, Fink G, Sherwood NM. Gonadotropin-releasing
hormone surge in proestrous rats. Nature 1976; 264: 461–463.
61 Roth C, Schricker M, Lakomek M, Strege AHI, Luft H, Munzel U, Wuttke
W, Jarry H. Autoregulation of the gonadotropin-releasing hormone
(GnRH) system during puberty: effects of antagonistic versus agonistic
GnRH analogs in a female rat model. J Endocrinol 2001; 169: 361–
371.
62 Zanisi M, Messi EM, Motta M, Martini L. Ultrashort feedback control of
luteinizing hormone-releasing hormone neuronal system. Endocrinology
1987; 121: 2199–2204.
63 Wu TJ, Pierotti AR, Jakubowski M, Sheward WJ, Glucksman MJ, Smith
AI, King JC, Fink G, Roberts JL. Endopeptidase EC 3.4.24.15 presence in
the rat median eminence and hypophysial portal blood and its modula-
tion of the luteinizing hormone surge. J Neuroendocrinol 1997; 9:
813–822.
64 Shrimpton CN, Glucksman MJ, Lew RA, Tullai JW, Margulies EH,
Roberts JL, Smith AI. Thiol activation of endopeptidase EC 3.4.24.15. A
novel mechanism for the regulation of catalytic activity. J Biol Chem
1997; 272: 17393–17399.
65 Smith AI, Tetaz T, Roberts JL, Glucksman MJ, Clarke IJ, Lew RA. The role
of EC 3.4.24.15 in the post-secretory regulation of peptide signals. Bio-
chimie 1994; 76: 288–294.
66 Roberts JL, Mani SK, Woller MJ, Glucksman MJ, Wu TJ. LHRH-(1-5): a
bioactive peptide regulating reproduction. Trends Endocrinol Metab
2007; 18: 386–392.
67 Wu TJ, Mani SK, Glucksman MJ, Roberts JL. Stimulation of luteinizing
hormone-releasing hormone (LHRH) gene expression in GT1-7 cells by
its metabolite, LHRH-(1-5). Endocrinology 2005; 146: 280–286.
68 Larco DO, Williams M, Schmidt L, Sabel N, Woller MJ, Wu TJ.
Autoshortloop feedback regulation of pulsatile gonadotropin-releasing
©2016 British Society for NeuroendocrinologyJournal of Neuroendocrinology, 2016, 28, 10.1111/jne.12405
Peripheral and central hormonal control of reproduction 9of12
hormone (GnRH) secretion by its metabolite, GnRH-(1-5). Endocr J
2015; 49: 470–478.
69 Wu TJ, Glucksman MJ, Roberts JL, Mani SK. Facilitation of lordosis in
rats by a metabolite of luteinizing hormone releasing hormone (LHRH).
Endocrinology 2006; 147: 2544–2549.
70 Cho-Clark MC, Larco DO, Semsarzadeh N, Vasta FC, Mani SK, Wu TJ.
GnRH-(1-5) transactivates EGFR in Ishikawa human endometrial cells via
an orphan G protein-coupled receptor. Mol Endocrinol 2014; 28: 80–98.
71 Larco DO, Cho-Clark M, Mani SK, Wu TJ. The metabolite GnRH-(1-5)
inhibits the migration of immortalized GnRH neurons. Endocrinology
2013; 154: 783–795.
72 Larco DO, Semsarzadeh N, Cho-Clark M, Mani SK, Wu TJ. b-Arrestin 2
is a mediator of GnRH-(1-5) signaling in immortalized GnRH neurons.
Endocrinology 2013; 154: 4726–4736.
73 Larco DO, Cho-Clark M, Semsarzadeh N, Mani SK, Wu TJ. The novel
actions of the metabolite GnRH-(1-5) are mediated by a G protein-
coupled receptor. Front Endocrinol 2013; 4: 83.
74 Regard JB, Sato IT, Coughlin SR. Anatomical profiling of G protein-
coupled receptor expression. Cell 2008; 135: 561–571.
75 Lee DK, George SR, Cheng R, Nguyen T, Liu Y, Brown M, Lynch KR,
O’Dowd BF. Identification of four novel G protein-coupled receptors
expressed in the brain. Brain Res Mol Brain Res 2001; 86: 13–22.
76 Bates B, Zhang L, Nawoschik S, Kodangattil S, Tseng E, Kopsco D,
Kramer A, Shan Q, Taylor N, Johnson J, Sun Y, Chen HM, Blatcher M,
Paulsen JE, Pausch MH. Characterization of Gpr101 expression and
G-protein coupling selectivity. Brain Res 2006; 1087: 1–14.
77 Cho-Clark M, Larco DO, Zahn BR, Mani SK, Wu TJ. GnRH-(1-5) activates
matrix metallopeptidase-9 to release epidermal growth factor and pro-
mote cellular invasion. Mol Cell Endocrinol 2015; 415: 114–125.
78 Clarke IJ, Sari IP, Qi Y, Smith JT, Parkington HC, Ubuka T, Iqbal J, Li
Q, Tilbrook A, Morgan K, Pawson AJ, Tsutsui K, Millar RP, Bentley
GE. Potent action of RFamide-related peptide-3 on pituitary gonado-
tropes indicative of a hypophysiotropic role in the negative regula-
tion of gonadotropin secretion. Endocrinology 2008; 149: 5811–
5821.
79 Bentley GE, Jensen JP, Kaur GJ, Wacker DW, Tsutsui K, Wingfield JC.
Rapid inhibition of female sexual behavior by gonadotropin-inhibitory
hormone (GnIH). Horm Behav 2006; 49: 550–555.
80 Tsutsui K, Saigoh E, Ukena K, Teranishi H, Fujisawa Y, Kikuchi M, Ishii
S, Sharp PJ. A novel avian hypothalamic peptide inhibiting gonadotro-
pin release. Biochem Biophys Res Commun 2000; 275: 661–667.
81 Ubuka T, Ukena K, Sharp PJ, Bentley GE, Tsutsui K. Gonadotropin-inhi-
bitory hormone inhibits gonadal development and maintenance by
decreasing gonadotropin synthesis and release in male quail.
Endocrinology 2006; 147: 1187–1194.
82 Ciccone NA, Dunn IC, Boswell T, Tsutsui K, Ubuka T, Ukena K, Sharp PJ.
Gonadotrophin inhibitory hormone depresses gonadotrophin alpha and
follicle-stimulating hormone beta subunit expression in the pituitary of
the domestic chicken. J Neuroendocrinol 2004; 16: 999–1006.
83 Ubuka T, Tsutsui K. Gonadotropin-inhibitory hormone inhibits aggres-
sive behavior of male quail by increasing neuroestrogen synthesis in
the brain beyond its optimum concentration. Gen Comp Endocrinol
2014; 205: 49–54.
84 Bentley GE, Perfito N, Ukena K, Tsutsui K, Wingfield JC. Gonadotropin-
inhibitory peptide in song sparrows (Melospiza melodia) in different
reproductive conditions, and in house sparrows (Passer domesticus)
relative to chicken-gonadotropin-releasing hormone. J Neuroendocrinol
2003; 15: 794–802.
85 Rizwan MZ, Porteous R, Herbison AE, Anderson GM. Cells expressing
RFamide-related peptide-1/3, the mammalian gonadotropin-inhibitory
hormone orthologs, are not hypophysiotropic neuroendocrine neurons
in the rat. Endocrinology 2009; 150: 1413–1420.
86 Anderson GM, Relf HL, Rizwan MZ, Evans JJ. Central and peripheral
effects of RFamide-related peptide-3 on luteinizing hormone and pro-
lactin secretion in rats. Endocrinology 2009; 150: 1834–1840.
87 Ubuka T, Kim S, Huang YC, Reid J, Jiang J, Osugi T, Chowdhury VS,
Tsutsui K, Bentley GE. Gonadotropin-inhibitory hormone neurons inter-
act directly with gonadotropin-releasing hormone-I and -II neurons in
European starling brain. Endocrinology 2008; 149: 268–278.
88 Kriegsfeld LJ, Mei DF, Bentley GE, Ubuka T, Mason AO, Inoue K, Ukena
K, Tsutsui K, Silver R. Identification and characterization of a gonado-
tropin-inhibitory system in the brains of mammals. Proc Natl Acad Sci
USA 2006; 103: 2410–2415.
89 Ukena K, Ubuka T, Tsutsui K. Distribution of a novel avian gonadotro-
pin-inhibitory hormone in the quail brain. Cell Tissue Res 2003; 312:
73–79.
90 Ubuka T, Morgan K, Pawson AJ, Osugi T, Chowdhury VS, Minakata H,
Tsutsui K, Millar RP, Bentley GE. Identification of human GnIH homo-
logs, RFRP-1 and RFRP-3, and the cognate receptor, GPR147 in the
human hypothalamic pituitary axis. PLoS ONE 2009; 4: e8400.
91 Ubuka T, Lai H, Kitani M, Suzuuchi A, Pham V, Cadigan PA, Wang A,
Chowdhury VS, Tsutsui K, Bentley GE. Gonadotropin-inhibitory hor-
mone identification, cDNA cloning, and distribution in rhesus macaque
brain. J Comp Neurol 2009; 517: 841–855.
92 Murakami M, Matsuzaki T, Iwasa T, Yasui T, Irahara M, Osugi T, Tsutsui
K. Hypophysiotropic role of RFamide-related peptide-3 in the inhibition
of LH secretion in female rats. J Endocrinol 2008; 199: 105–112.
93 Bentley GE, Kriegsfeld LJ, Osugi T, Ukena K, O’Brien S, Perfito N, Moore
IT, Tsutsui K, Wingfield JC. Interactions of gonadotropin-releasing hor-
mone (GnRH) and gonadotropin-inhibitory hormone (GnIH) in birds
and mammals. J Exp Zool A Comp Exp Biol 2006; 305: 807–814.
94 Smith JT, Coolen LM, Kriegsfeld LJ, Sari IP, Jaafarzadehshirazi MR, Mal-
tby M, Bateman K, Goodman RL, Tilbrook AJ, Ubuka T, Bentley GE,
Clarke IJ, Lehman MN. Variation in kisspeptin and RFamide-related
peptide (RFRP) expression and terminal connections to gonadotropin-
releasing hormone neurons in the brain: a novel medium for seasonal
breeding in the sheep. Endocrinology 2008; 149: 5770–5782.
95 Mason AO, Duffy S, Zhao S, Ubuka T, Bentley GE, Tsutsui K, Silver R,
Kriegsfeld LJ. Photoperiod and reproductive condition are associated
with changes in RFamide-related peptide (RFRP) expression in Syrian
hamsters (Mesocricetus auratus). J Biol Rhythms 2010; 25: 176–185.
96 Piekarski DJ, Jarjisian SG, Perez L, Ahmad H, Dhawan N, Zucker I,
Kriegsfeld LJ. Effects of pinealectomy and short day lengths on repro-
duction and neuronal RFRP-3, Kisspeptin, and GnRH in female Turkish
hamsters. J Biol Rhythms 2014; 29: 181–191.
97 Ubuka T, Inoue K, Fukuda Y, Mizuno T, Ukena K, Kriegsfeld LJ, Tsutsui K.
Identification, expression, and physiological functions of Siberian ham-
ster gonadotropin-inhibitory hormone. Endocrinology 2012; 153: 373–
385.
98 Ubuka T, Bentley GE, Ukena K, Wingfield JC, Tsutsui K. Melatonin
induces the expression of gonadotropin-inhibitory hormone in the
avian brain. Proc Natl Acad Sci USA 2005; 102: 3052–3057.
99 Chowdhury VS, Yamamoto K, Ubuka T, Bentley GE, Hattori A, Tsutsui K.
Melatonin stimulates the release of gonadotropin-inhibitory hormone
by the avian hypothalamus. Endocrinology 2010; 151: 271–280.
100 Calisi RM, Diaz-Munoz SL, Wingfield JC, Bentley GE. Social and breed-
ing status are associated with the expression of GnIH. Genes Brain
Behav 2011; 10: 557–564.
101 Calisi RM, Rizzo NO, Bentley GE. Seasonal differences in hypothalamic EGR-
1 and GnIH expression following capture-handling stress in house sparrows
(Passer domesticus). Gen Comp Endocrinol 2008; 157: 283–287.
102 McGuire NL, Koh A, Bentley GE. The direct response of the gonads to
cues of stress in a temperate songbird species is season-dependent.
PeerJ 2013; 1: e139.
©2016 British Society for Neuroendocrinology Journal of Neuroendocrinology, 2016, 28, 10.1111/jne.12405
10 of 12 L. M. Rudolph et al.
103 McGuire NL, Kangas K, Bentley GE. Effects of melatonin on peripheral
reproductive function: regulation of testicular GnIH and testosterone.
Endocrinology 2011; 152: 3461–3470.
104 McGuire NL, Bentley GE. Neuropeptides in the gonads: from evolution
to pharmacology. Front Pharmacol 2010; 1: 114.
105 McGuire NL, Bentley GE. A functional neuropeptide system in verte-
brate gonads: gonadotropin-inhibitory hormone and its receptor in
testes of field-caught house sparrow (Passer domesticus). Gen Comp
Endocrinol 2010; 166: 565–572.
106 Bentley GE, Ubuka T, McGuire NL, Chowdhury VS, Morita Y, Yano T,
Hasunuma I, Binns M, Wingfield JC, Tsutsui K. Gonadotropin-inhibitory
hormone and its receptor in the avian reproductive system. Gen Comp
Endocrinol 2008; 156: 34–43.
107 Zhao S, Zhu E, Yang C, Bentley GE, Tsutsui K, Kriegsfeld LJ. RFamide-
related peptide and messenger ribonucleic acid expression in mam-
malian testis: association with the spermatogenic cycle. Endocrinology
2010; 151: 617–627.
108 Oishi H, Klausen C, Bentley GE, Osugi T, Tsutsui K, Gilks CB, Yano T,
Leung PC. The human gonadotropin-inhibitory hormone ortholog RFa-
mide-related peptide-3 suppresses gonadotropin-induced progesterone
production in human granulosa cells. Endocrinology 2012; 153: 3435–
3445.
109 Kirby ED, Geraghty AC, Ubuka T, Bentley GE, Kaufer D. Stress increases
putative gonadotropin inhibitory hormone and decreases luteinizing
hormone in male rats. Proc Natl Acad Sci USA 2009; 106: 11324–
11329.
110 Geraghty AC, Muroy SE, Zhao S, Bentley GE, Kriegsfeld LJ, Kaufer D.
Knockdown of hypothalamic RFRP3 prevents chronic stress-induced
infertility and embryo resorption. Elife 2015; 4: e04316. doi: 10.7554/
eLife.04316.
111 Dickens MJ, Bentley GE. Stress, captivity, and reproduction in a wild
bird species. Horm Behav 2014; 66: 685–693.
112 Lynn SE, Perfito N, Guardado D, Bentley GE. Food, stress, and circulat-
ing testosterone: cue integration by the testes, not the brain, in male
zebra finches (Taeniopygia guttata). Gen Comp Endocrinol 2015; 215:
1–9.
113 Ernst DK, Lynn SE, Bentley GE. Differential response of GnIH in the
brain and gonads following acute stress in a songbird. Gen Comp
Endocrinol 2016; 227: 51–57.
114 Lopes PC, Wingfield JC, Bentley GE. Lipopolysaccharide injection induces
rapid decrease of hypothalamic GnRH mRNA and peptide, but does not
affect GnIH in zebra finches. Horm Behav 2012; 62: 173–179.
115 Simoni M, Gromoll J, Nieschlag E. The follicle-stimulating hormone
receptor: biochemistry, molecular biology, physiology, and pathophysi-
ology. Endocr Rev 1997; 18: 739–773.
116 Dufau ML. The luteinizing hormone receptor. Annu Rev Physiol 1998;
60: 461–496.
117 Tena-Sempere M. Explorin the role of ghrelin as a novel regulator of
gonadal function. Growth Horm IGF Res 2005; 15: 83–88.
118 Mayerhofer A. Leydig cell regulation by catecholamines and neuroen-
docrine messengers. In: Payne AH, Hardy MP, Russell LD, eds. The Ley-
dig Cell. St Louis, MO: Cache River Press, 1996; 407–417.
119 McGuire NI, Bentley GE. Neuropeptides in the gonads: from evolution
to pharmacology. Front Pharmacol 2010; 1: 1–13.
120 Teerds KJ, Dorrington JH. Localization of transforming growth factor
beta1 and beta2 during testicular development in the rat. Biol Reprod
1993; 48: 40–45.
121 Wang X, Shen CL, Dyson MT, Eimerl S, Orly J, Hutson JC, Stocco DM.
Cyclooxygenase-2 regulation of the age-related decline in testosterone
biosynthesis. Endocrinology 2005; 146: 4202–4208.
122 Tinajero JC, Fabri A, Ciocca DR, Dufau ML. Serotonin secretion from rat
Leydig cells. Endocrinology 1993; 133: 3026–3029.
123 Valenti S, Thellung S, Florio T, Guisti M, Schettini G, Giordano G. A
novel mechanism for the melatonin inhibition of testosterone secretion
by rat Leydig cells: reduction of GnRH-induced increase in cytosolic
Ca2+.J Mol Endocrinol 1999; 23: 299–306.
124 Dufau ML, Tinajero JC, Fabbri A. Corticotropin-releasing factor: an
antireproductive hormone of the testis. FASEB J 1993; 7: 299–307.
125 Frungieri MB, Mayerhofer A, Zitta K, Pignataro OP, Calandra RS. Direct
effect of melatonin on Syrian hamster testes: melatonin subtype 1a
receptors, inhibition of androgen production, and interaction with the
local corticotropin-releasing hormone system. Endocrinology 2005;
146: 1541–1552.
126 Frungieri MB, Gonzalez-Calvar SI, Rubio M, Ozu M, Lustig L, Calandra
RS. Serotonin in golden hamster testes: testicular levels, immunolocal-
ization and role during sexual development and photoperiodic regres-
sion-recrudescence transition. Neuroendocrinology 1999; 69: 299–308.
127 Frungieri MB, Zitta K, Pignataro OP, Gonzalez-Calvar SI, Calandra RS.
Interactions between testicular serotoninergic, catecholaminergic, and
corticotropin-releasing hormone systems modulating cAMP and testos-
terone production in the golden hamster. Neuroendocrinology 2002;
76: 35–46.
128 Rossi SP, Matzkin ME, Terradas C, Ponzio R, Puigdomenech E, Levalle
O, Calandra RS, Frungieri MB. New insights into melatonin/CRH signal-
ing in hamster Leydig cells. Gen Comp Endocrinol 2012; 178: 153–
163.
129 Carrillo-Vico A, Calvo JR, Abreu P, Lardone PJ, Garc
ıa-Mauri~
no S,
Reiter RJ, Guerrero JM. Evidence of melatonin synthesis by human
lymphocytes and its physiological significance: possible role as intra-
crine, autocrine, and/or paracrine substance. FASEB J 2004; 18: 537–
539.
130 Rossi SP, Windschuettl S, Matzkin ME, Terradas C, Ponzio R, Puig-
domenech E, Levalle O, Calandra RS, Mayerhofer A, Frungieri MB.
Melatonin in testes of infertile men: evidence for anti-proliferative and
anti-oxidant effects on local macrophage and mast cell populations.
Andrology 2014; 2: 436–449.
131 Frungieri MB, Calandra RS, Mayerhofer A, Matzkin ME. Cyclooxygenase
and prostaglandins in somatic cell populations of the testes. Reproduc-
tion 2015; 149: R169–R180.
132 Matzkin ME, Pelizzari EH, Rossi SP, Calandra RS, Cigorraga SB, Frungieri
MB. Exploring the cyclooxygenase 2 (COX2)/15d-D(12,14)PGJ(2) system
in hamster Sertoli cells: regulation by FSH/testosterone and relevance
to glucose uptake. Gen Comp Endocrinol 2012; 179: 254–264.
133 Frungieri MB, Gonzalez-Calvar SI, Parborell F, Albrecht M, Mayerhofer
A, Calandra RS. Cyclooxygenase-2 and prostaglandin F2 alpha in Syrian
hamster Leydig cells: inhibitory role on luteinizing hormone/human
chorionic gonadotropin-stimulated testosterone production. Endocrinol-
ogy 2006; 147: 4476–4485.
134 Schell C, Frungieri MB, Albrecht M, Gonzalez-Calvar SI, K
ohn FM,
Calandra RS, Mayerhofer A. A prostaglandin D2 system in the human
testis. Fertil Steril 2007; 88: 233–236.
135 Brugh VM, Matschke HM, Lipschultz LI. Male factor infertility. Endocri-
nol Metab Clin North Am 2003; 32: 689–707.
136 Wulff C, Wilson H, Largue P, Duncan WC, Armstrong DG, Fraser HM.
Angiogenesis in the human corpus luteum: localization and changes in
angiopoietins, tie-2, and vascular endothelial growth factor messenger
ribonucleic acid. J Clin Endocrinol Metab 2000; 85: 4302–4309.
137 Abramovich D, Parborell F, Tesone M. Effect of a vascular endothelial
growth factor (VEGF) inhibitory treatment on the folliculogenesis and
ovarian apoptosis in gonadotropin-treated prepubertal rats. Biol Reprod
2006; 75: 434–441.
138 Parborell F, Abramovich D, Tesone M. Intrabursal administration of the
antiangiopoietin 1 antibody produces a delay in rat follicular develop-
ment associated with an increase in ovarian apoptosis mediated by
©2016 British Society for NeuroendocrinologyJournal of Neuroendocrinology, 2016, 28, 10.1111/jne.12405
Peripheral and central hormonal control of reproduction 11 of 12
changes in the expression of BCL2 related genes. Biol Reprod 2008;
78: 506–513.
139 Abramovich D, Irusta G, Parborell F, Tesone M. Intrabursal injection of
vascular endothelial growth factor trap in ECG-treated prepubertal rats
inhibits proliferation and increases apoptosis of follicular cells involving
the PI3K/AKT signaling pathway. Fertil Steril 2010; 93: 1369–1377.
140 Irusta G, Abramovich D, Parborell F, Tesone M. Direct survival role of
vascular endothelial growth factor (VEGF) on rat ovarian follicular cells.
Mol Cell Endocrinol 2010; 325: 93–100.
141 Hernandez F, Peluffo MC, Stouffer RL, Irusta G, Tesone M. Role of the
DLL4-NOTCH system in PGF2alpha-induced luteolysis in the pregnant
rat. Biol Reprod 2011; 84: 859–865.
142 Kopan R, Ilagan MX. The canonical notch signaling pathway: unfolding
the activation mechanism. Cell 2009; 137: 216–233.
143 Accialini P, Hernandez SF, Bas D, Pazos MC, Irusta G, Abramovich D,
Tesone M. A link between notch and progesterone maintains the func-
tionality of the rat corpus luteum. Reproduction 2015; 149: 1–10.
144 Pinilla L, Aguilar E, Dieguez C, Millar RP, Tena-Sempere M. Kisspeptins
and reproduction: physiological roles and regulatory mechanisms.
Physiol Rev 2012; 92: 1235–1316.
145 Vikman J, Ahr
en BD. Inhibitory effect of kisspeptins on insulin secre-
tion from isolated mouse islets. Diabetes Obes Metab 2009; Suppl 4:
197–201.
146 Castellano JM, Gaytan M, Roa J, Vigo E, Navarro VM, Bellido C, Dieguez
C, Aguilar E, Sanchez-Criado JE, Pellicer A, Pinilla L, Gaytan F, Tena-
Sempere M. Expression of KiSS-1 in rat ovary: putative local regulator
of ovulation? Endocrinology 2006; 147: 4852–4862.
147 Ricu M, Ramirez VD, Paredes AH, Lara HE. Evidence for a celiac gan-
glion-ovarian kisspeptin neural network in the rat: intraovarian anti-
kisspeptin delays vaginal opening and alters estrous cyclicity.
Endocrinology 2012; 153: 4966–4977.
148 Laoharatchatathanin T, Terashima R, Yonezawa T, Kurusu S, Kawami-
nami M. Augmentation of Metastin/Kisspeptin mRNA expression by the
proestrous luteinizing hormone surge in granulosa cells of rats: impli-
cations for luteinization. Biol Reprod 2015; 93: 15.
149 Fernandois D, Na E, Cuevas F, Cruz G, Lara HE, Paredes AH. Kisspeptin
is involved in ovarian follicular development during the fertility and
subfertility periods of rats. J Endocrinol 2016; 228: 161–170.
150 Gayt
an F, Gayt
an M, Castellano JM, Romero M, Roa J, Aparicio B, Gar-
rido N, S
anchez-Criado JE, Millar RP, Pellicer A, Fraser HM, Tena-Sem-
pere M. KiSS-1 in the mammalian ovary: distribution of kisspeptin in
human and marmoset and alterations in KiSS-1 mRNA levels in a rat
model of ovulatory dysfunction. Am J Physiol Endocrinol Metab 2009;
296: 520–531.
151 Xiao Y, Ni Y, Huang Y, Wu J, Grossmann R, Zhao R. Effects of kisspep-
tin-10 on progesterone secretion in cultured chicken ovarian granulosa
cells from preovulatory (F1-F3) follicles. Peptides 2011; 32: 2091–2097.
152 Peng J, Tang M, Zhang BP, Zhang P, Zhong T, Zong T, Yang B, Kuang
HB. Kisspeptin stimulates progesterone secretion via the Erk1/2 mito-
gen-activated protein kinase signaling pathway in rat luteal cells. Fertil
Steril 2013; 99: 1436–1443.
153 Cedars MI. Biomarkers of ovarian reserve–do they predict somatic
aging? Semin Reprod Med 2013; 31: 443–451.
154 Hoyer PB. Damage to ovarian development and function. Cell Tissue
Res 2005; 322: 99–106.
155 Lenton EA, Deketser DM, Woodward AJ. Inhibin concentrations
throughout the menstrual cycle of normal, infertile and older women
compared with those during spontaneous conceptions cycles. J Clin
Endocrinol Metab 1991; 73: 1180–1186.
156 Semenova II. Adrenergic innervation of ovaries in Stein-Leventhal syn-
drome. Vestn Akad Med Nauk SSSR 1969; 24: 58–62.
157 Heider U, Pedal I, Spanel-Borowski K. Increase in nerve fibers and loss
of mast cells in polycystic and postmenopausal ovaries. Fertil Steril
2001; 75: 1141–1147.
158 Acu~
na E, Fornes R, Fernandois D, Garrido MP, Greiner M, Lara HE,
Paredes AH. Increases in norepinephrine release and ovarian cyst for-
mation during ageing in the rat. Reprod Biol Endocrinol 2009; 7: 64.
159 Chavez-Genaro R, Lombide P, Dominguez R, Rosas P, Vazquez-Cuevas
F. Sympathetic pharmacological denervation in ageing rats: effects on
ovulatory response and follicular population. Reprod Fertil Dev 2007;
19: 954–960.
160 Cooke BM, Woolley CS. Sexually dimorphic synaptic organization of the
medial amygdala. J Neurosci 2005; 25: 10759–10767.
161 Matsumoto A, Arai Y. Development of sexual dimorphism in synaptic
organization in the ventromedial nucleus of the hypothlamus in rats.
Neurosci Lett 1986; 68: 165–168.
162 Christensen A, Dewing P, Micevych P. Membrane-initiated estradiol sig-
naling induces spinogenesis required for female sexual receptivity. J
Neurosci 2011; 31: 17583–17589.
163 Akama KT, Thompson LI, Milner TA, McEwen BS. Post-synaptic density-
95 (PSD-95) binding capacity of G-protein-coupled receptor 30
(GPR30), an estrogen receptor that can be identified in hippocampal
dendritic spines. J Biol Chem 2013; 288: 6438–6450.
164 Waters EM, Thompson LI, Patel P, Gonzales AD, Ye HZ, Filardo EJ, Clegg
DJ, Gorecka J, Akama KT, McEwen BS, Milner TA. G-protein-coupled
estrogen receptor 1 is anatomically positioned to modulate synaptic
plasticity in the mouse hippocampus. J Neurosci 2015; 35: 2384–2397.
165 Garcia BL, Mana A, Kim A, Sinchak K. Antagonism of Estrogen Recep-
tors Facilitates Sexual Receptivity through Opioid Circuits in the Arcu-
ate Nucleus of the Hypothalamus and the Medial Preoptic Nucleus in
Estradiol Primed Non-Receptive Female Rats. San Diego, CA: Society
for Neuroscience, 2010.
166 Kuo J, Hamid N, Bondar G, Prossnitz E, Micevych P. Membrane estrogen
receptors stimulate intracellular calcium release and progesterone syn-
thesis in hypothalamic astrocytes. J Neurosci 2010; 30: 12950–12957.
167 Qiu J, Ronnekleiv O, Kelly M. Modulation of hypothalamic neuronal
activity through a novel G-protein-coupled estrogen membrane recep-
tor. Steroids 2008; 73: 985–991.
168 Qui J, Bosch M, Tobias S, Grandy D, Scanlan T, Ronnekleiv O, Kelly M.
Rapid signaling of estrogen in hypothalamic neurons involves a novel
G-protein-coupled estrogen receptor that activates protein kinase C. J
Neurosci 2003; 23: 9529–9540.
169 Christensen A, Micevych P. A novel membrane estrogen receptor activated
by STX induces female sexual receptivity through an interacation with
mGluR1a. Neuroendocrinology 2013; 97: 363–368.
170 Nishida H, Okabe S. Direct astrocytic contacts regulate local maturation
of dendritic spines. J Neurosci 2007; 27: 331–340.
171 Liu T, Kong D, Shah BP, Ye C, Koda S, Saunders A, Ding JB, TYang Z,
Sabatini BL, Lowell BB. Fasting activation of AgRP neurons requires
NMDA receptors and involves spinogenesis and increased excitatory
tone. Neuron 2012; 73: 511–522.
172 Mills RH, Sohn RK, Micevych PE. Estrogen-induced mu-opioid receptor
internalization in the medial preoptic nucleus is mediated via neu-
ropeptide Y-Y1 receptor activation in the arcuate nucleus of female
rats. J Neurosci 2004; 24: 947–955.
173 Kasai H, Matsuzaki M, Noguchi J, Yasumatsu N, Nakahara H. Struc-
ture-stability-function relationships of dendritic spines. Trends Neurosci
2003; 26: 360–368.
©2016 British Society for Neuroendocrinology Journal of Neuroendocrinology, 2016, 28, 10.1111/jne.12405
12 of 12 L. M. Rudolph et al.
... Those neurons are the main node of reproduction in animals [24] as they act on the gonadotrophs in the anterior pituitary to trigger luteinizing hormone (LH) and follicle-stimulating hormone release into the circulation. Those glycoproteins will act at the level of the gonads to stimulate sex steroid production and release to enable reproduction [24][25][26]. Thus, since aggressive behavior itself [27,28] as well as successive aggressive interactions and social isolation are known to increase plasma testosterone [8,14,17,18,27,28] and to some extent LH [27,29] (main hormonal output of GnRH neurons), one could hypothesize that GnRH neurons are recruited during aggressive interactions and might be impacted by the social context. ...
Article
Full-text available
Although the participation of sex hormones and sex hormone-responsive neurons in aggressive behavior has been extensively studied, the role of other systems within the hypothalamus-pituitary-gonadal (HPG) axis remains elusive. Here we assessed how the gonadotropin-releasing hormone (GnRH) and kisspeptin systems are impacted by escalated aggression in male mice. We used a combination of social isolation and aggression training (IST) to exacerbate mice's aggressive behavior. Next, low-aggressive (group-housed, GH) and highly aggressive (IST) mice were compared regarding neuronal activity in the target populations and hormonal levels, using immunohistochemistry and ELISA, respectively. Finally, we used pharmacological and viral approaches to manipulate neuropeptide signaling and expression, subsequently evaluating its effects on behavior. IST mice exhibited enhanced aggressive behavior compared to GH controls, which was accompanied by elevated neuronal activity in GnRH neurons and arcuate nucleus kisspeptin neurons. Remarkably, IST mice presented an increased number of kisspeptin neurons in the anteroventral periventricular nucleus (AVPV). In addition, IST mice exhibited elevated levels of luteinizing hormone (LH) in serum. Accordingly, activation and blockade of GnRH receptors (GnRHR) exacerbated and reduced aggression, respectively. Surprisingly, kisspeptin had intricate effects on aggression, i.e., viral ablation of AVPV-kisspeptin neurons impaired the training-induced rise in aggressive behavior whereas kisspeptin itself strongly reduced aggression in IST mice. Our results indicate that IST enhances aggressive behavior in male mice by exacerbating HPG-axis activity. Particularly, increased GnRH neuron activity and GnRHR signaling were found to underlie aggression whereas the relationship with kisspeptin remains puzzling. Neuropsychopharmacology; https://doi.
... In addition, COX-2 was found to be implicated in testicular inflammation related to idiopathic infertility in biopsies of men with impaired spermatogenesis [18]. Alteration of PGs synthesis in the testis is another potential reason for idiopathic male fertility [19]. However, both COX-1, and COX-2 isoforms were not found in normal human testes [20]. ...
Article
Full-text available
Arachidonic acid (AA) is a polyunsaturated fatty acid that is involved in male fertility. Human seminal fluid contains different prostaglandins: PGE (PGE1 and PGE2), PGF2α, and their specific 19-hydroxy derivatives, 18,19-dehydro derivatives of PGE1 and PGE2. The objective of this study is to synthesize the available literature of in vivo animal studies and human clinical trials on the association between the AA pathway and male fertility. PGE is significantly decreased in the semen of infertile men, suggesting the potential for exploitation of PGE agonists to improve male fertility. Indeed, ibuprofen can affect male fertility by promoting alterations in sperm function and standard semen parameters. The results showed that targeting the AA pathways could be an attractive strategy for the treatment of male fertility.
... Only little proteolytic activity was detected in Sertoli cell-conditioned supernatants, suggesting that the proteases under investigation were mostly found on Sertoli cell membranes [23]. The peptide hormones utilized in this research were discovered to be involved in the endocrine, paracrine, or autocrine control of testicular cells [24]. As a result, the membrane-associated proteases described here might be implicated in the metabolism and deactivation of these peptides [25,26]. ...
Article
Full-text available
Non-obstructive azoospermia (NOA) is a serious cause of male infertility. The Sertoli cell responds to androgens and takes on roles supporting spermatogenesis, which may cause infertility. This work aims to enhance the genetic diagnosis of NOA via the discovery of new and hub genes implicated in human NOA and to better assess the odds of successful sperm extraction according to the individual’s genotype. Whole exome sequencing (WES) was done on three NOA patients to find key genes involved in NOA. We evaluated genome-wide transcripts (about 50,000 transcripts) by microarray between the Sertoli of non-obstructive azoospermia and normal cells. The microarray analysis of three human cases with different non-obstructive azoospermia revealed that 32 genes were upregulated, and the expressions of 113 genes were downregulated versus the normal case. For this purpose, Enrich Shiny GO, STRING, and Cytoscape online evaluations were applied to predict the functional and molecular interactions of proteins and then recognize the master pathways. The functional enrichment analysis demonstrated that the biological process (BP) terms “inositol lipid-mediated signaling”, “positive regulation of transcription by RNA polymerase II”, and “positive regulation of DNA-templated transcription” significantly changed in upregulated differentially expressed genes (DEGs). The BP investigation of downregulated DEGs highlighted “mitotic cytokinesis”, “regulation of protein-containing complex assembly”, “cytoskeleton-dependent cytokinesis”, and the “peptide metabolic process”. Overrepresented molecular function (MF) terms in upregulated DEGs included “ubiquitin-specific protease binding”, “protease binding”, “phosphatidylinositol trisphosphate phosphatase activity”, and “clathrin light chain binding”. Interestingly, the MF analysis of the downregulated DEGs revealed overexpression in “ATPase inhibitor activity”, “glutathione transferase activity”, and “ATPase regulator activity”. Our findings suggest that these genes and their interacting hub proteins could help determine the pathophysiologies of germ cell abnormalities and infertility.
... Thus, the AHA could be significant in combining stress and energy regulation. The periventricular hypothalamic nucleus is known to release, somatostatin and growth hormone, exhibiting sex differences in LH secretion [128], possibly due to differential ERβ expression between the sexes [129]. Thus, the periventricular hypothalamic nucleus is a major contributor to reproduction and gonadal functions. ...
Article
Full-text available
Estrogens are among important contributing factors to many sex differences in neuroendocrine regulation of energy homeostasis induced by stress. Research in this field is warranted since chronic stress-related psychiatric and metabolic disturbances continue to be top health concerns, and sex differences are witnessed in these aspects. For example, chronic stress disrupts energy homeostasis, leading to negative consequences in the regulation of emotion and metabolism. Females are known to be more vulnerable to the psychological consequences of stress, such as depression and anxiety, whereas males are more vulnerable to the metabolic consequences of stress. Sex differences that exist in the susceptibility to various stress-induced disorders have led researchers to hypothesize that gonadal hormones are regulatory factors that should be considered in stress studies. Further, estrogens are heavily recognized for their protective effects on metabolic dysregulation, such as anti-obesogenic and glucose-sensing effects. Perturbations to energy homeostasis using laboratory rodents, such as physiological stress or over-/under- feeding dietary regimen prevalent in today’s society, offer hints to the underlying mechanisms of estrogenic actions. Metabolic effects of estrogens primarily work through estrogen receptor α (ERα), which is differentially expressed between the sexes in hypothalamic nuclei regulating energy metabolism and in extrahypothalamic limbic regions that are not typically associated with energy homeostasis. In this review, we discuss estrogenic actions implicated in stress-induced sex-distinct metabolic disorders.
... Given the complicated and dynamic sequence of events occurring during spermatogenesis (mitosis, meiosis and cell differentiation), with control systems required at both central and peripheral levels [74,75], a copious amount of ROS is physiologically generated by germ cells as by-products of their metabolism [76]. However, a moderate quantity of ROS is also convenient for regular functions, such as cell signalling, homeostasis, sperm capacitation, and sperm-egg interaction [77][78][79]. ...
Article
Full-text available
Besides ATP production, mitochondria are key organelles in several cellular functions, such as steroid hormone biosynthesis, calcium homoeostasis, intrinsic apoptotic pathway, and the generation of reactive oxygen species (ROS). Despite the loss of the majority of the cytoplasm occurring during spermiogenesis, mammalian sperm preserves a number of mitochondria that rearrange in a tubular structure at the level of the sperm flagellum midpiece. Although sperm mitochondria are destroyed inside the zygote, the integrity and the functionality of these organelles seem to be critical for fertilization and embryo development. The aim of this review was to discuss the impact of mitochondria-produced ROS at multiple levels in sperm: the genome, proteome, lipidome, epigenome. How diet, aging and environmental pollution may affect sperm quality and offspring health—by exacerbating oxidative stress—will be also described.
... FSH stimulate the spermatogenesis while LH have role in the production and release of testosterone, male reproductive hormone, from the leydig cells essential for spermatogenesis and for reproductive system development like testis and prostate. 66 It also regulate the expression of secondary sexual characteristic in males like increased muscle mass, bone mass, the endocrine system and growth of body-hair. 67 FSH function is to stimulate the process of spermatogenesis while LH play role in stimulating the testosterone synthesis and release. ...
Article
Full-text available
Current study was planned to explore the therapeutic response of different doses of hydroethanolic extract of Epimedium grand-iflorum leaves in male albino rats. Phytochemical analysis, HPLC and FTIR spectroscopy results revealed the presence of wide range of phenolic compounds and functional groups, respectively. Further, extract not induced significant hemolysis (7.56 + 1.297%) against PBS (3.65 + 0.35%) as negative control; while have significant clot lysis (44 + 5.2%) potential, exhibited DPPH (78.87 + 5.427%) scavenging, H 2 O 2 (31.82 + 3.491%) scavenging, antioxidant and reducing power activities. In vivo experimentation in albino male rats' revealed that administration of different doses (50, 100, 200 mg/Kg b.w.) of extract orally for 42 days after CCl 4 intoxication significantly (P < 0.05) restore the selected parameters including liver enzymes, renal profiles, and stress markers and significantly (P < 0.05) increased reproductive hormones like testosterone, luteinizing hormone, follicle stimulating hormone and prolactin while significantly (P < 0.05) decreased progesterone and estradiol toward normal in dose dependent manner. Significant (P < 0.05) improvement in the structural architecture of testicular tissue particularly in high dose group (200 mg/Kg b.w.) was also observed. Results revealed E. grandiflorum has significant therapeutic response to address the healthcare problems particularly of impotency.
... Recently, PNX-14 has been shown to be a ligand of the orphan Rhodopsin receptor G protein-coupled receptor 173 (GPR173). GPR173 is a 42-kDa seven-transmembrane receptor with a sequenced mRNA of 1122 bases [20,21]. However, the effects of GPR173 and the extent of its expression are poorly understood. ...
Article
Full-text available
IntroductionNeuroinflammation is a key aspect of various injuries and diseases of the central nervous system and brain, including stroke, Alzheimer’s, Parkinson’s, multiple sclerosis, etc. Phoenixin-14 is a naturally occurring pleiotropic peptide involved in reproduction, anxiety, pain, and other functions.Materials and Methods Primary astrocytes were isolated from new-born pups of c57bl/6 mice. The gene expression of GPR173, CHOP, and GADD34 was measured by real-time PCR. Protein expression was assessed by western blot analysis. Secretions of IL-1β and IL-18 were determined by ELISA.ResultsPhoenixin-14 (PNX-14) is a ligand for the G protein-coupled receptor GPR173, which we demonstrate to be expressed in astrocytes and suppressed by exposure to lipopolysaccharide (LPS). Endoplasmic reticulum (ER) stress resulting from injury or disease leads to the unfolded protein response, which is mediated by the activation of transcription factors including eIF-2α, ATF4, and CHOP, and regulated by GADD34. ER stress also leads to a robust neuroinflammatory response, which is mediated by HMGB1-induced activation of the NLRP3 inflammasome and subsequent production of IL-1β and IL-18. In the present study, we demonstrate that PNX-14 could attenuate LPS-induced ER stress response and NLRP3 inflammasome activation in mouse cerebral astrocytes. Our findings show that PNX-14 could suppress the production of ROS as well as the decrease in SOD induced by LPS. PNX-14 also inhibited HMGB1-mediated NLRP3 inflammasome activation and production of IL-1β and IL-18. Through a GPR173 siRNA knockdown experiment, we further demonstrate that GPR173 knockdown abolished the effects of PNX-14 on LPS-induced NLRP3 expression and IL-18 production.Conclusion These findings suggest that PNX-14 may have potential in the treatment of neuroinflammation.
Chapter
Reproduction is an important biological process of species evolution, it is leading to new offspring from parents. The pituitary gland (the master gland) in the endocrine system in coordination with the hypothalamus plays a vital role in the reproductive system, differentiation, and different physiological functions in the entire stages of life and its circadian rhythm in both males and females. The chapter deals with the prominent role played by the brain, endocrine system, and gonads axis through complex communicating signals. The male gonads are the location of testicles, interstitial tissue (Leydig cells), and peritubular myoid cells. Sertoli cells act as “nurse and stem” cells, spermatogenesis, spermiogenesis were explained. Gonad’s steroid hormones (Androgens) characteristics are specified in male reproductive activity, biological actions along with the regulator hormones (follicle-stimulating hormone, luteinizing hormone, and hypothalamic–pituitary–Leydig cell axis. Ovaries are female reproductive glands. The chapter describes the tissue zones of ovaries, puberty, and two main functions (exocrine and endocrine) controlled and coordinated by the hypothalamus and the pituitary. Female sex hormones in pre-puberty (Estrogen) and post-puberty (estradiol, estrone, progesterone, and inhibin), and sources (ovarian follicle and corpus luteum) were discussed in detail. Structure of steroid hormones discussed with the role of the endometrium, regulation of ovarian functions, and puberty by endocrine and immune system. The different phases of the ovarian cycle are explained to regulate gonadotropins, follicular growth, steroid synthesis, non-functional corpus albicans (infertile ovum), and regulation of the ovarian cycle (pre-ovulatory phase, ovulation phase, and post-ovulatory phase). The involvement of estradiol metabolites, enzyme aromatase, the hormone oxytocin, matrix metalloprotease, cytokines, and vasoconstriction in corpus luteum formation is related along with maintenance and regression. Synthesis of ovarian hormones (β-estradiol, estrone, and estriol) after puberty discussed with important functions of progesterone, regulatory roles of inhibin, activin, and follistatin in the physiology of testis and ovary. Dehydroepiandrosterone (DHEA) involvement was discussed for menopause in elderly women and andropause in men. The chapter declares that the classic theory of cessation of oocytes production after birth was canceled. Both human neonatal and adult ovarian germline stem-cell precursors (ovarian surface cells) have the capability for oogenic/differentiating and producing functional oocytes, so it renews the oocyte pool (neo-oogenesis) and ensures renewal during the prime reproductive period, with follicular cooperation under the regulation of the endocrine, immune systems, and cellular support. After the prime reproductive period, aging starts, and menopause occurs because of the immunoregulatory changes that cause cessation and terminate neo-oogenesis and follicular renewal in vivo despite the existence of germline stem cell precursors. The rest of the oocytes in the primordial follicles retain ovarian function but advancing age (aging oocytes) correlates positively with the occurrence of fetal chromosomal abnormality. This chapter discusses the topics related to regulation of male and female reproductive functions.KeywordsFollicle-stimulating hormoneLuteinizing hormoneOvaryTestisReproductive hormoneStem cell
Chapter
The pituitary is considered the master gland of the organism since it controls a multiplicity of biological processes, including growth, reproduction, whole-body metabolism, or stress. This gland is comprised by two main structurally and functionally distinct areas named adenohypophysis and neurohypophysis. These regions have different developmental origins and exert diverse functional roles in whole human physiology through the secretion of a variety of hormones, including growth hormone (GH), prolactin (PRL), luteinizing hormone (LH), follicle-stimulating hormone (FSH), thyrotropin-stimulating hormone (TSH), adrenocorticotropic hormone (ACTH), melanocyte-stimulating hormone (MSH), oxytocin, and vasopressin. In this context, the secretion of pituitary hormones is a complex physiological process finely regulated by a plethora of signals. Initially, it was thought that the synthesis and release of pituitary hormones was mainly controlled by central signals. However, it is now known that each pituitary cell type has a particular profile of receptors for a wide variety of central and peripheral neuroendocrine signals, which are intracellularly integrated to finely regulate the secretion of all the types of pituitary hormones. This chapter will comprehensively review the most relevant information regarding the physiology and regulation of the secretion of the different pituitary hormones.
Article
Adverse effects of drugs on male reproductive system can be categorized as pre-testicular, testicular, and post-testicular. Pre-testicular adverse effects disrupt the hypothalamic–pituitary–gonadal (HPG) axis, generally by interfering with endocrine function. It is known that the HPG axis has roles in the maintenance of spermatogenesis and sexual function. The hypothalamus secretes gonadotropin-releasing hormone (GnRH) which enters the hypophyseal portal system to stimulate the anterior pituitary. The anterior pituitary secretes gonadotropins, follicle-stimulating hormone (FSH) and luteinizing hormone (LH) which are vital for spermatogenesis, into the blood. The FSH stimulates the Sertoli cells for the production of regulatory molecules and nutrients needed for the maintenance of spermatogenesis, while the LH stimulates the Leydig cells to produce and secrete testosterone. Many neurotransmitters influence the hypothalamic-pituitary regulation, consequently the HPG axis, and can consequently affect spermatogenesis and sexual function. Psychotropic drugs including antipsychotics, antidepressants, and mood stabilizers that all commonly modulate dopamine, serotonin, and GABA, can affect male spermatogenesis and sexual function by impairment of the hypothalamic-pituitary regulation, act like endocrine-disrupting chemicals. Otherwise, studies have shown the relationship between decreased sperm quality and psychotropic drugs treatment. Therefore, it is important to investigate the adverse reproductive effects of psychotropic drugs which are frequently used during reproductive ages in males and to determine the role of the hypothalamic-pituitary regulation axis on possible pathologies.
Article
Full-text available
Whereas it is well established that chronic stress induces female reproductive dysfunction, whether stress negatively impacts fertility and fecundity when applied prior to mating and pregnancy has not been explored. In this study, we show that stress that concludes 4 days prior to mating results in persistent and marked reproductive dysfunction, with fewer successful copulation events, fewer pregnancies in those that successfully mated, and increased embryo resorption. Chronic stress exposure led to elevated expression of the hypothalamic inhibitory peptide, RFamide-related peptide-3 (RFRP3), in regularly cycling females. Remarkably, genetic silencing of RFRP3 during stress using an inducible-targeted shRNA completely alleviates stress-induced infertility in female rats, resulting in mating and pregnancy success rates indistinguishable from non-stress controls. We show that chronic stress has long-term effects on pregnancy success, even post-stressor, that are mediated by RFRP3. This points to RFRP3 as a potential clinically relevant single target for stress-induced infertility.
Article
Full-text available
We have previously reported that kisspeptin (KP) may be under the control of the sympathetic innervation of the ovary. Considering that the sympathetic activity of the ovary increases with ageing, it is possible that ovarian KP also increases during this period and participates in follicular development. To evaluate this possibility, we determined ovarian KP expression and its action on follicular development during reproductive ageing in rats. We measured ovarian KP mRNA and protein levels in 6-, 8-, 10- and 12-month-old rats. To evaluate follicular developmental changes, intraovarian administration of KP or its antagonist, peptide 234 (P234), was performed using a mini-osmotic pump, and to evaluate FSH receptor (FSHR) changes in the senescent ovary, we stimulated cultured ovaries with KP, P234 and isoproterenol (ISO). Our results show that KP expression in the ovary was increased in 10- and 12-month-old rats compared with 6-month-old rats, and this increase in KP was strongly correlated with the increase in ovarian norepinephrine observed with ageing. The administration of KP produced an increase in corpora lutea and type III follicles in 6- and 10-month-old rats, which was reversed by P234 administration at 10 months. In addition, KP decreased the number and size of antral follicles in 6- and 10-month-old rats, while P234 administration produced an increase in these structures at the same ages. In ovarian cultures KP prevented the induction of FSHR by isoproterenol. These results suggest that intraovarian KP negatively participates in the acquisition of FSHR, indicating a local role in the regulation of follicular development and ovulation during reproductive ageing.
Article
Full-text available
Kisspeptin, encoded by Kiss1, stimulates GnRH neurons to govern reproduction. In rodents, estrogen-sensitive kisspeptin neurons in the AVPV/PeN are thought to mediate sex steroid-induced positive feedback induction of the preovulatory LH surge. These kisspeptin neurons coexpress estrogen and progesterone receptors and display enhanced neuronal activation during the LH surge. However, while estrogen regulation of kisspeptin neurons has been well-studied, the role of progesterone signaling in regulating kisspeptin neurons is unknown. Here, we tested whether progesterone action specifically in kisspeptin cells is essential for proper LH surge and fertility. We used Cre-lox technology to generate transgenic mice lacking progesterone receptors exclusively in kisspeptin cells (termed KissPRKOs). Male KissPRKOs displayed normal fertility and gonadotropin levels. In stark contrast, female KissPRKOs displayed earlier puberty onset and significant impairments in fertility, evidenced by fewer births and substantially reduced litter size. KissPRKOs also had fewer ovarian corpora lutea, suggesting impaired ovulation. To ascertain if this reflects a defect in the ability to generate sex steroid-induced LH surges, females were exposed to an estradiol positive feedback paradigm. Unlike control females which displayed robust LH surges, KissPRKO females did not generate notable LH surges and expressed significantly blunted cfos induction in AVPV/PeN kisspeptin neurons, indicating that progesterone receptor signaling in kisspeptin neurons is required for normal kisspeptin neuronal activation and LH surges during positive feedback. Our novel findings demonstrate that progesterone signaling specifically in kisspeptin cells is essential for the positive feedback induction of normal LH surges and ovulation, and hence, for normal fertility in females.
Article
Classically, 17beta-estradiol (E2) is thought to control homeostatic functions such as reproduction, stress responses, feeding, sleep cycles, temperature regulation, and motivated behaviors through transcriptional events. Although it is increasingly evident that E2 can also rapidly activate kinase pathways to have multiple downstream actions in CNS neurons, the receptor(s) and the signal transduction pathways involved have not been identified. We discovered that E2 can alter mu-opioid and GABA neurotransmission rapidly through nontranscriptional events in hypothalamic GABA, proopiomelanocortin (POMC), and dopamine neurons. Therefore, we examined the effects of E2 in these neurons using whole-cell recording techniques in ovariectomized female guinea pigs. E2 reduced rapidly the potency of the GABAB receptor agonist baclofen to activate G-protein-coupled, inwardly rectifying K+ channels in hypothalamic neurons. These effects were mimicked by the membrane impermeant E2-BSA and selective estrogen receptor modulators, including a new diphenylacrylamide compound, STX, that does not bind to intracellular estrogen receptors alpha or beta, suggesting that E2 acts through a unique membrane receptor. We characterized the coupling of this estrogen receptor to a Galpha(q)-mediated activation of phospholipase C, leading to the upregulation of protein kinase Cdelta and protein kinase A activity in these neurons. Moreover, using single-cell reverse transcription-PCR, we identified the critical transcripts, PKCdelta and its downstream target adenylyl cyclase VII, for rapid, novel signaling of E2 in GABA, POMC, and dopamine neurons. Therefore, this unique Gq-coupled estrogen receptor may be involved in rapid signaling in hypothalamic neurons that are critical for normal homeostatic functions.
Article
The corpus luteum (CL) is a transient endocrine gland in which angiogenesis and vessel regression occur during cyclic ovarian function in adult females. The CL differentiates from the follicle wall after ovulation by tissue remodeling and extensive neo-vascularization. Moreover, the integrity and function of the luteal vasculature decline during regression of the CL near the end of the cycle. The Notch family pathway, particularly Delta-like ligand 4 (Dll4) and its receptors Notch1 and 4 were recently identified as novel factors involved in angiogenesis. This system regulates cell fate decisions, including proliferation and death. The transmembrane receptors are cleaved upon binding of their ligands, leading to the release of the intracellular domain, which translocates to the nucleus where it functions as a transcriptional coactivator of regulatory genes of cellular fate. This processing requires the activity of two proteases, namely tumour necrosis factor alpha-converting enzyme (TACE) and presenilin/gamma-secretase. Given the implication for angiogenesis in the CL lifespan, we investigated the role of the Notch pathway in luteal function and regression of the CL during pregnancy in rats. To determine if the Notch signaling pathway is implicated in CL regression during pregnancy, pregnant rats were injected with a luteolytic dose of PGF2alpha (400µg, ip) or vehicle on day 19 of pregnancy. Ovaries were removed 4 and 24 hours after the treatment and CL's collected by microdissection. Total RNA was isolated and protein extraction was performed for real-time PCR and western blot assay respectively, of Dll4, Notch1 and 4. Dll4 mRNA levels significantly decreased at 4 hours after PGF2alpha administration, whereas no changes were detected at the protein level. In contrast, both mRNA and protein expression of Notch1 and 4 receptors significantly decreased 4 hours after PGF2alpha administration. However, by 24 hours after treatment there were no further changes in the expression of the ligand and receptors. To elucidate if the Notch system regulates luteal function during pregnancy, either vehicle (control) or the gamma-secretase inhibitor DAPT (10 µg/ovary) was injected into the bursa of both ovaries on day 19 of pregnancy. Twenty-four hours after treatment, blood samples and CLs were collected. Luteal function was evaluated by measuring serum progesterone (P4) by RIA and apoptosis was examined by DNA fragmentation in agarose gels. The levels of P4 significantly decreased after DAPT administration. However, apoptotic DNA fragmentation in the CL was not affected by DAPT treatment. These results suggest that the Notch signaling pathway promotes CL function and is acutely suppressed during PGF2alpha-induced CL regression in pregnant rats. Supported by: ANPCYT (BID 1201 OC-AR PICT 99:05-06384), NIH-FIRCA RO3-TW007041 and NIH-NCCR RR00163. (poster)
Article
In rat Leydig cells, serotonin (5HT) binds to 5HT2 receptors and stimulates the secretion of CRF which in turn acts as an inhibitor of gonadotropin-induced cAMP formation and androgen production. In the present study we defined the regulation of 5HT secretion in cultured Leydig cells. Adult Leydig cells secreted considerable quantities of 5HT (100-150 pg/10(6) cells per 10 min). The release of 5HT was acutely stimulated by hCG (ED50, 1.1 pM) with maximal stimulation at 10 pM hCG (160%). Forskolin also increased (+220%) 5HT release from cultures (ED50, 50 nM) while TPA was much less effective (+20%), indicating a major role for cAMP in gonadotropin-induced 5HT release. This was confirmed by the finding that 8-Br cAMP (1 mM) was an effective stimulus of 5HT release (+360%). Similar increases of 5HT release by hCG were observed in the absence of extracellular Ca2+. However, ionomycin was a potent stimulus of 5HT release, indicating that elevation of cytoplasmic [Ca2+] could also induce amine secretion. The 5HT content of Leydig cells ranged from 300 to 350 pg/10(6) cells, and decreased during stimulation of 5HT release. Also, immunohistochemical studies revealed specific staining of 5 HT in interstitial cells of the adult rat testis. These studies demonstrated that rat Leydig cells contain and secrete 5HT, and that 5HT release is stimulated by gonadotropin acting primarily through a cAMP-mediated mechanism.
Article
In the extracellular space, the gonadotropin-releasing hormone (GnRH) is metabolized by the zinc metalloendopeptidase EC3.4.24.15 (EP24.15) to form the pentapeptide, GnRH-(1-5). GnRH-(1-5) diverges in function and mechanism of action from GnRH in the brain and periphery. GnRH-(1-5) acts on the orphan G protein-coupled receptor 101 (GPR101) to sequentially stimulate epidermal growth factor (EGF) release, phosphorylate the EGF receptor (EGFR), and facilitate cellular migration. These GnRH-(1-5) actions are dependent on matrix metallopeptidase (MMP) activity. Here, we demonstrated that these GnRH-(1-5) effects are dependent on increased MMP-9 enzymatic activity in the Ishikawa and ECC-1 cell lines. Furthermore, the effects of GnRH-(1-5) mediated by GPR101 and the subsequent increase in MMP-9 enzymatic activity lead to an increase in cellular invasion. These results suggest that GnRH-(1-5) and GPR101 regulation of MMP-9 may have physiological relevance in the metastatic potential of endometrial cancer cells. Copyright © 2015. Published by Elsevier Ireland Ltd.