Next Article in Journal
Multiplication of ampC upon Exposure to a Beta-Lactam Antibiotic Results in a Transferable Transposon in Escherichia coli
Next Article in Special Issue
Impairment of the Hypothalamus–Pituitary–Thyroid Axis Caused by Naturally Occurring GATA2 Mutations In Vitro
Previous Article in Journal
SARS-CoV-2 Impairs Dendritic Cells and Regulates DC-SIGN Gene Expression in Tissues
Previous Article in Special Issue
The Role of Histone Deacetylase 3 Complex in Nuclear Hormone Receptor Action
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Kisspeptin Neurons and Estrogen–Estrogen Receptor α Signaling: Unraveling the Mystery of Steroid Feedback System Regulating Mammalian Reproduction

1
Laboratory of Animal Reproduction, Graduate School of Bioagricultural Sciences, Nagoya University, Nagoya 464-8601, Japan
2
Faculty of Veterinary Medicine, Okayama University of Science, Imabari 794-8555, Japan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(17), 9229; https://doi.org/10.3390/ijms22179229
Submission received: 29 July 2021 / Revised: 21 August 2021 / Accepted: 23 August 2021 / Published: 26 August 2021
(This article belongs to the Special Issue Steroids and Lipophilic Hormones, and Their Actions 2.0)

Abstract

:
Estrogen produced by ovarian follicles plays a key role in the central mechanisms controlling reproduction via regulation of gonadotropin-releasing hormone (GnRH) release by its negative and positive feedback actions in female mammals. It has been well accepted that estrogen receptor α (ERα) mediates both estrogen feedback actions, but precise targets had remained as a mystery for decades. Ever since the discovery of kisspeptin neurons as afferent ERα-expressing neurons to govern GnRH neurons, the mechanisms mediating estrogen feedback are gradually being unraveled. The present article overviews the role of kisspeptin neurons in the arcuate nucleus (ARC), which are considered to drive pulsatile GnRH/gonadotropin release and folliculogenesis, in mediating the estrogen negative feedback action, and the role of kisspeptin neurons located in the anteroventral periventricular nucleus-periventricular nucleus (AVPV-PeN), which are thought to drive GnRH/luteinizing hormone (LH) surge and consequent ovulation, in mediating the estrogen positive feedback action. This implication has been confirmed by the studies showing that estrogen-bound ERα down- and up-regulates kisspeptin gene (Kiss1) expression in the ARC and AVPV-PeN kisspeptin neurons, respectively. The article also provides the molecular and epigenetic mechanisms regulating Kiss1 expression in kisspeptin neurons by estrogen. Further, afferent ERα-expressing neurons that may regulate kisspeptin release are discussed.

1. Introduction

It has been well accepted that estrogen produced by the ovary plays an indispensable role in the female reproductive system via its feedback actions on gonadotropin-releasing hormone (GnRH) release in mammals. The central mechanisms of the estrogen feedback actions on GnRH release have been a mystery for decades. This is because no report has been available to show the expression of estrogen receptor α (ERα), a critical receptor isoform required for estrogen feedback actions, in the hypothalamic GnRH neurons. Intensive studies on the hypothalamic kisspeptin neurons, which express ERα, have been gradually unraveling the central mechanisms of the negative and positive feedback actions of estrogen on GnRH release. In this article, the physiological significance of the estrogen negative and positive feedback actions on the tonic pulsatile and surge modes of GnRH release—which control folliculogenesis and ovulation in female mammals, respectively—is outlined. Further, the molecular and epigenetic mechanisms mediating the regulation of kisspeptin gene (Kiss1) expression by estrogen–ERα signaling and afferent ERα-expressing neurons that may mediate estrogen-dependent modulation of kisspeptin release from the hypothalamic kisspeptin neurons are also discussed.

2. Feedback Actions of Estrogen on Pulsatile and Surge-Modes of Gonadotropin-Releasing Hormone (GnRH)/Gonadotropin Release

Mammalian reproduction is orchestrated by the interaction of hormones secreted by the hypothalamus–pituitary–gonadal (HPG) axis. Estrogen secreted from the ovary, downstream of the axis, feeds back to the higher hierarchy hypothalamus and pituitary to regulate GnRH/gonadotropin release. Estrogen production is stimulated by the tonic pulsatile release of gonadotropins, such as luteinizing hormone (LH) and follicle-stimulating hormone (FSH), from the anterior pituitary gland, under the control of GnRH pulses. During the follicular development, circulating estrogen fine-tunes pulsatile release of GnRH to keep circulating levels of LH and FSH adequately. This estrogen action is referred to as “the negative feedback action of estrogen” on GnRH pulses. Under appropriate stimulation by LH and FSH, ovarian follicles develop into a large and mature state. Estrogen production and release gradually increase along with the follicular development, and consequent high levels of circulating estrogen derived from mature follicles (also known as Graafian follicles), in turn, induce a large release of hypothalamic GnRH and then pituitary LH (GnRH/LH surge). This is the so-called “positive feedback action of estrogen” on GnRH release, and the LH surge consequently evokes ovulation. Therefore, the circulating levels of estrogen serve as a messenger for transmitting the maturity status of ovarian follicles to the hypothalamus, which plays a pinnacle role in the hierarchical control of the HPG axis in female mammals.
The presence of hypothalamic GnRH was predicted by Harris and Jacobsohn in the early 1950s by showing that the function of the pituitary graft was restored only when the graft was placed under the median eminence in hypophysectomized rats [1]. This notion was further validated by McCann and colleagues in the early 1960s by showing the LH-releasing ability of hypothalamic extracts in rats [2,3]. The tonic pulsatile LH release, the preovulatory LH surge, and their regulation by estrogen feedback are then clearly demonstrated by Knobil and colleagues from the late 1960s to the early 1970s [4,5,6,7,8]. They demonstrated that tonic LH release was found in most periods of the menstrual cycle, and LH surge was found only in the midcycle before ovulation in humans and rhesus monkeys [4,5]. It was also demonstrated that ovariectomy increased plasma LH concentration, indicating the negative feedback action of some ovarian humoral factor(s) on tonic pulsatile LH release in rhesus monkeys [6]. Importantly, estrogen replacement at a physiological level (not a preovulatory level) suppressed tonic pulsatile LH release in ovariectomized (OVX) rhesus monkeys [8]. These findings suggest that estrogen secreted from the ovary is a major humoral factor that exerts its negative feedback action on the tonic pulsatile LH release. On the other hand, an administration of estrogen at a preovulatory dose induces a vast release of LH similar to the spontaneous preovulatory LH surge, suggesting that the high dose of estrogen exerts its positive feedback action on the LH surge-generating system [7,8]. In 1971, GnRH was isolated from the hypothalamus of pigs and sheep by two groups led by Schally and Guillemin [9,10]. In the early 1990s, Moenter et al. [11,12,13] suggested that GnRH dominantly controls LH release by showing that GnRH pulses and GnRH surge in the pituitary portal circulation were synchronized with LH pulses and LH surge in the peripheral circulation, respectively, in sheep. In addition, Pau et al. [14] showed simultaneous GnRH and LH surges in rhesus monkeys as well. To date, the pulsatile and surge modes of GnRH release have been hypothesized to be driven by the independent hypothalamic mechanisms, so-called “GnRH pulse and surge generators”, respectively [15,16,17]. It is plausible that estrogen regulates the activity of GnRH pulse and surge generators via the negative and positive feedback actions, respectively.

3. Indispensable Role of Estrogen Receptor α for Mammalian Reproduction

Accumulating evidence indicates that ERα is a critical estrogen receptor isoform responsible for both the negative and positive feedback actions of estrogen on GnRH/LH release. In fact, Esr1 (coding ERα) knockout mice [18,19] and rats [20] show hypersecretion of both LH and estrogen, indicating that ERα mainly mediates the estrogen negative feedback action. Further, the Esr1 knockout mice and rats fail to show ovulation, albeit enlarged cystic follicles are found in Esr1 knockout mice and rats. This suggests that ERα is also responsible for the estrogen positive feedback action [18,19,20]. On the other hand, reproductive function of Esr2 (coding ERβ, known as another ER) knockout animal models were reportedly varied between animal models: Esr2 knockout mice showed normosecretion of LH and estrogen [19] and are subfertile with a small litter size [21,22]; Esr2 knockout rats are infertile with a lack of LH surge and ovulation [23]. In addition, previous studies demonstrated that selective antagonism of estrogen–ERα signaling, but not estrogen–ERβ signaling, eliminated the endogenous LH surge in rats [24]. Taken together, these findings suggest that ERα mainly mediates both estrogen feedback actions on GnRH/LH release.
In general, the ERα is known as a ligand-activated transcriptional factor that activates or represses the expression of target genes. The estrogen-bound ERα is reported to bind to the estrogen response element (ERE) in the target genes to control gene expression [25]. In addition, it is suggested that the estrogen-bound ERα interacts with other transcription factors, such as AP-1 and NF-κB, and the complex controls target gene expression via binding to non-ERE response elements through the transcriptional partner [26,27,28]. Intriguingly, a previous study demonstrated that ERα knock-in/knockout (KIKO) mice, in which a mutant ERα (E207A/G208A) lacks the binding ability for the ERE but is capable of interacting with other transcriptional partners [29,30], showed the negative, but not the positive, feedback action of estrogen on LH release [31]. These findings suggest that the negative feedback action of estrogen on GnRH pulses is likely mediated via some gene(s) controlled by the ERE-independent estrogen–ERα signaling and that the estrogen positive feedback action is likely mediated via some gene(s) controlled by the ERE-dependent estrogen–ERα signaling.

4. Possible Targets of the Negative and Positive Feedback Action of Estrogen in the Brain

Precise targets of the negative and positive feedback actions of estrogen on the GnRH pulse and surge generation have been a mystery for many years of the 20th century because no report has been available to show ERα expression in GnRH neurons [32]. Therefore, the most plausible explanations are that certain hypothalamic ERα-expressing cells serve as targets of the negative and positive feedback actions of estrogen on GnRH pulse and surge generation and that such ERα-expressing cells transmit the estrogen signals to GnRH neurons. A large number of ERα-expressing cells were found in the several hypothalamic nuclei—such as the anteroventral periventricular nucleus (AVPV), preoptic area (POA), arcuate nucleus (ARC), and ventromedial nucleus (VMH)—at both the mRNA and protein levels, as well as the paraventricular nucleus (PVN) and suprachiasmatic nucleus (SCN), where ERα expression was evident at the mRNA level in rats [33,34]. Similarly, ERα was mainly found in the POA, ARC, and VMH at both the mRNA and protein levels and in the PVN at the mRNA level in sheep [35,36]. These findings were well consistent with previous studies showing that radiolabeled estrogen was accumulated in the POA, ARC, and VMH in rats [37].
Previous studies suggest that the ARC is one of the most possible targets of negative feedback action of estrogen: Smith and Davidson [38] showed that estrogen implants in the mediobasal hypothalamus (MBH), including the ARC suppressed plasma LH levels in OVX rats in 1974; Akema et al. [39] showed that estrogen implants in the ARC suppressed LH pulses in OVX rats in 1983; furthermore, Nagatani et al. [40] showed that estrogen micro-implants in the ARC suppressed LH pulses in both fasted and re-fed OVX rats, while the estrogen implants in either the PVN or brainstem A2 region suppressed LH pulses in only fasted rats in 1994. These findings suggested that the negative feedback action of estrogen may be mediated by ERα-expressing neurons located in the ARC under the normal nutritional condition and by multiple ERα-expressing neurons located in the ARC, PVN, and brainstem A2 region under the malnutritional condition. The negative feedback actions of estrogen under the malnutritional condition are likely mediated by de novo synthesized ERα in the PVN and brainstem A2 region because 48 h fasting increases the number of ERα-immunoreactive cells in the PVN and brainstem A2 region in OVX rats [41].
Previous studies suggest that ERα-expressing neurons in the AVPV and/or POA are the most possible targets of estrogen positive feedback action: Kawakami et al. [42] and Goodman [43] demonstrated in the late 1970s that estrogen implants into the AVPV or neighboring POA induced the LH surge in OVX rats; Wiegand et al. [44,45] showed in the late 1980s that an electrolytic lesion around the AVPV abolished the estrogen-induced LH surge in OVX rats; Petersen et al. [46,47] demonstrated in the late 1980s that implants of estrogen antagonists, such as LY-10074 or keoxifene, in the AVPV-POA region prevented estrogen-induced LH surge in OVX rats. These reports suggest that the ERα-expressing cells in the AVPV-POA region serve as targets of the estrogen positive feedback actions to induce GnRH/LH surge.

5. Kisspeptin Neurons as Targets of the Negative and Positive Feedback Actions of Estrogen

Intensive studies during the past 20 years demonstrate that ERα expression is evident in the hypothalamic kisspeptin neurons in rodents [48,49,50,51] and sheep [52] and that kisspeptin serves as a potent secretagogue of gonadotropin release in rodents [50,53,54,55,56,57,58], ruminants [59,60], and primates [61,62]. To date, it is well accepted that ERα-expressing kisspeptin neurons mainly mediate the estrogen feedback on GnRH release in mammals, and the possible mechanism mediating the feedback effect is discussed in detail later in this article.
Kisspeptin was first discovered as an endogenous ligand for GPR54, an orphan Gq-coupled G-protein coupled receptor (GPCR), in humans in 2001 [63,64]. In 2003, two independent research groups reported that inactivating mutations in the GPR54 gene caused hypogonadotropic hypogonadism in humans [65,66]. These important findings shed light on the fact that kisspeptin–GPR54 signaling plays a pivotal role in the brain mechanism controlling GnRH/gonadotropin release and then puberty and fertility in mammals [65,66]. As expected, inactivating mutations in the KISS1 gene (coding kisspeptin) also caused hypogonadotropic hypogonadism in humans [67]. The infertile phenotype in humans carrying inactivating mutations of the KISS1 or GPR54 genes was recapitulated in Kiss1 or Gpr54 knockout rodent models [66,68,69,70,71,72]. Importantly, GPR54 expression is evident in GnRH neurons in rodents [54,68,73,74,75], suggesting that kisspeptin directly stimulates GnRH release. Further, Kiss1 knockout rats show undetectable levels of LH and FSH even after ovariectomy, indicating failure of tonic pulsatile LH release [72]. In addition, the Kiss1 knockout rats also fail to show estrogen-induced LH surge. These findings suggest that kisspeptin–GPR54 signaling is indispensable for both GnRH pulse and surge generation and mediate feedback actions of estrogen on GnRH/LH release.
Histological studies in rodents revealed that cell bodies of kisspeptin neurons are mainly located in the anterior hypothalamic areas, such as the AVPV–periventricular nucleus continuum (AVPV-PeN), and in the posterior hypothalamic region—that is, the ARC [48,49,50,51,76,77,78]. Importantly, ERα was found in both populations of hypothalamic kisspeptin neurons, and Kiss1 expression is controlled by estrogen in a brain region-specific fashion in rodents [48,49,50,51,76]. More specifically, the ARC Kiss1 expression level was high at diestrus and was suppressed by estrogen treatment [48,49,50,51,76], whereas the AVPV-PeN Kiss1 expression level was high at the afternoon of proestrus and was increased by estrogen treatment in rodents [48,49,51,76]. These findings suggest that the ARC kisspeptin neurons are a target of the negative feedback action of estrogen and that the AVPV-PeN kisspeptin neurons are a target of the positive feedback action of estrogen. Figure 1 depicts the brain mechanism mediating the estrogen negative and positive feedback actions on GnRH/gonadotropin release to regulate follicular development and ovulation in rodents. As shown in the figure, it is most likely that the ARC kisspeptin neurons control GnRH/LH pulses via mediating the estrogen negative feedback action and that the AVPV-PeN kisspeptin neurons control GnRH/LH surge via mediating the estrogen positive feedback action.

5.1. The Molecular and Epigenetic Mechanism Mediating the Regulation of Arcuate Kiss1 Expression by Estrogen and the Role of arcuate Kisspeptin Neurons as the GnRH Pulse Generator in Mammals

To date, the ARC kisspeptin neurons have been considered to serve as a target of estrogen negative feedback action on GnRH pulse generation, and a similar population of kisspeptin neurons have been identified in the ARC in other species or infundibular nucleus in primates (equivalent to the ARC in others) of several mammalian species, including humans [79,80], macaque monkeys [80,81,82,83], sheep [52,84,85,86], goats [59,87,88], cattle [89], horse [90], pigs [91], and musk shrews [92]. Our studies and other previous studies demonstrated that estrogen treatment largely repressed ARC Kiss1 expression in rodents [48,49,50,51]. Similar to the rodent models, previous studies demonstrated estrogen-dependent repression of KISS1 expression in the ARC kisspeptin neurons in sheep [93,94] and the infundibular nucleus in primates including humans [80]. These findings suggest that the estrogen negative feedback action on ARC kisspeptin neurons would be largely common among mammalian species.
According to the studies with rodent models, estrogen-dependent repression of Kiss1 expression in ARC kisspeptin neurons is likely mediated via the ERE-independent pathway because estrogen repressed the ARC Kiss1 expression even in ERα KIKO mice [95,96,97]. In addition, our previous chromatin immunoprecipitation (ChIP) assay with antibodies against ERα and acetylated histone H3 revealed that estrogen-bound ERα induced histone H3 deacetylation of the Kiss1 promoter region in the ARC kisspeptin neurons by showing that estrogen treatment lowered acetylated histone H3 levels in the Kiss1 promoter region in mouse ARC tissue [98]. These findings suggest that an estrogen-dependent inactivating modification of histone H3 of the Kiss1 promoter region resulted in the repression of Kiss1 expression. Furthermore, our in vivo reporter assay utilizing Kiss1-GFP reporter mice suggested that the 5′-intergenic region of the Kiss1 gene is required for an induction of Kiss1 mRNA expression in the ARC of female mice [99]. Indeed, reporter mice carrying the 5′-truncated Kiss1-GFP transgene (RBRC09415 and RBRC09416) failed to display the GFP expression in ARC kisspeptin neurons even after ovariectomy. Importantly, the reporter mice displayed the GFP expression in the AVPV-PeN kisspeptin neurons in the presence of estrogen [99]. Furthermore, other reporter mice carrying the full-length of Kiss1-GFP transgene (RBRC09413) displayed the GFP expression in both ARC and AVPV-PeN kisspeptin neurons in OVX and estrogen-treated OVX conditions, respectively [99]. Taken together, we speculate that the estrogen-bound ERα may cancel interaction, which is most likely chromatin loop formation, between the Kiss1 promoter and 5′-intergenic enhancer regions, resulting in the repression of Kiss1 expression in ARC kisspeptin neurons even after the ovariectomy.
Collectively, we envisage the molecular mechanism of estrogen negative feedback action on ARC Kiss1 expression as shown in Figure 2. Briefly, circulating estrogen most likely binds to ERα in the ARC kisspeptin neurons, and then the estrogen-bound ERα coupled with unknown transcriptional partner(s) may repress Kiss1 expression via a non-classic ERE-independent pathway in ARC kisspeptin neurons. The estrogen-bound ERα may induce histone H3 deacetylation at the Kiss1 promoter, and the estrogen-bound ERα and/or this inactivating histone modification may unwind chromatin loops between the Kiss1 promoter and the 5′-intergenic regions of Kiss1 locus, resulting in the repression of ARC Kiss1 expression in ARC kisspeptin neurons.
The vast majority of ARC kisspeptin neurons reportedly express neurokinin B (NKB) and dynorphin A (Dyn), thus the ARC kisspeptin neurons are also called KNDy neurons [86,87,100,101,102]. Accumulating evidence suggests that the ARC KNDy neurons can serve as an intrinsic source of the GnRH pulse generator [103,104,105,106,107]. The notion was recently confirmed by our study showing that rescuing Kiss1 expression only in ARC Tac3 (NKB gene)-expressing neurons recovered LH pulses and follicular development in global Kiss1 knockout rats [108]. The multiple-unit activity (MUA) recording demonstrated that rhythmic increases in the MUA volley detected from the recording electrodes placed in close vicinity to ARC kisspeptin (KNDy) neurons were synchronized with LH pulses in goats [59,87]. In addition, conditional ARC-specific Kiss1 knockout by using the Cre-loxP system severely or partially suppressed LH pulses in rats [108] and mice [109,110] according to the knockout rates in each individual. Further, the fiber photometry recording revealed that the mouse ARC kisspeptin neurons displayed rhythmic increases in intracellular Ca2+ levels that correspond to LH pulses [111,112]. Thus, kisspeptin neurons may secrete kisspeptin in a pulsatile fashion and then induce GnRH/gonadotropin pulses. Indeed, Keen et al. [113] and Kurian et al. [114] showed pulsatile kisspeptin release that mostly corresponds to GnRH pulses at the median eminence in rhesus monkeys. Thus, the negative feedback action of estrogen directly acts on the intrinsic source of the GnRH pulse generator—namely, ARC kisspeptin neurons—and then suppresses GnRH/LH pulses. In this context, the profound suppression of GnRH/LH pulses before the afternoon LH surge in the female rodents in the presence of a high dose of estrogen may be due to the abovementioned epigenetic repression of ARC Kiss1 expression and consequent deficiency of kisspeptin in ARC kisspeptin neurons. Indeed, chronic treatment of preovulatory levels of estrogen profoundly suppresses tonic LH release in the morning (before LH surge) in OVX rats, and the estrogen treatment largely decreased Kiss1 expression as well as kisspeptin-immunoreactivity in the ARC of female rats [51].
In addition to the direct inhibiting action of estrogen on Kiss1 expression, estrogen may also inhibit the pulsatile activity of ARC kisspeptin neurons via other intra-kisspeptin neuronal mechanisms or some afferent ERα-expressing neurons to ARC kisspeptin neurons. The frequency of KNDy neuronal activity recorded by the MUA volley was increased and decreased by a central administration of NKB and Dyn, respectively, in goats [87,104]. A majority of KNDy neurons reportedly express both tachykinin NK3 receptor, a Gq-coupled GPCR for NKB, and kappa-opioid receptor (KOR), a Gi-coupled GPCR for Dyn in mice [102,115,116,117], rats [118,119], and sheep [120,121]. Considering the stimulatory or inhibitory signaling of NKB or Dyn, respectively, these findings suggest that the pulsatile activity of ARC kisspeptin (KNDy) neurons is controlled by NKB and Dyn in an autocrine/paracrine manner (please see review articles for details, [103,104,105]). Previous studies showed that estrogen decreased NKB gene (Tac2 in mice and Tac3/TAC3 in other mammals) expression in the ARC of mice [96,97,122] and sheep [123] and in the infundibular nucleus of rhesus monkeys [124]. In addition, estrogen decreased Dyn gene (Pdyn) expression in the ARC of mice [95,96] and rats [125]. These results suggest that estrogen may modulate kisspeptin release from the ARC KNDy neurons via changing stimulatory NKB and inhibitory Dyn inputs to the KNDy neurons.
Interestingly, the proestrous level of estrogen repressed ARC Kiss1 expression [50,51], whereas the diestrous level of estrogen, which exerted negative feedback action of LH pulses [78], failed to suppress ARC Kiss1 expression in female rats [50,51]. The dose of estrogen required for the repression of ARC Kiss1 expression raises the possibility that certain afferent ERα-expressing neurons to ARC kisspeptin neurons may be involved in the negative feedback action of estrogen on kisspeptin release from the ARC kisspeptin neurons. This notion is supported by a previous study showing that estrogen effectively decreased plasma LH concentration even in kisspeptin neuron-specific ERα knockout mice, whose ARC Kiss1 expression was not repressed by estrogen treatment [126]. One of the candidates mediating the estrogen negative feedback action would be Dyn neurons located in the PVN. Our studies showed that estrogen increased the number of Pdyn-expressing cells in the PVN [118] and that nor-binaltorphimine (nor-BNI), a KOR antagonist, enhanced LH pulses in estrogen-treated OVX rats but not in OVX rats without estrogen replacement [127]. Further, our recent study showed that glucoprivation suppressed LH pulses and induced fos (coding c-Fos, a marker of neuronal activation) expression in PVN Dyn neurons, while central KOR antagonism blocked the glucoprivic suppression of LH pulses in estrogen-treated OVX rats [118]. These findings suggest that PVN Dyn neurons may partly mediate estrogen negative feedback action to suppress kisspeptin release via KOR expressed in ARC kisspeptin neurons and then suppress pulsatile GnRH/LH release.

5.2. The Role of Anteroventral Periventricular Nucleus-Periventricular Nucleus (AVPV-PeN)/Preoptic Area (POA) Kisspeptin Neurons as the GnRH/Luteinizing Hormone (LH) Surge Generator and the Molecular and Epigenetic Mechanism Mediating the Regulation of AVPV-PeN/POA Kiss1 Expression by Estrogen Positive Feedback Action

The AVPV-PeN kisspeptin neurons have been considered to serve as a target of estrogen positive feedback action on GnRH surge generation in rodents, as already mentioned in the article. To date, kisspeptin neurons were found in the POA in several mammalian species, including macaque monkeys [81,83], sheep [85], goats [88], cattle [89], and musk shrews [92], as well as in the PeN in pigs [91]. Previous studies demonstrated that estrogen treatment largely increased AVPV-PeN Kiss1 expression [48,49,51,128] and induced c-Fos expression in AVPV-PeN kisspeptin neurons in OVX rodent models [49,51]. Similarly, our and other previous studies demonstrated estrogen-induced KISS1 and/or c-Fos expression in the POA/PeN kisspeptin neurons of macaque monkeys [81,83], sheep [85], goats [88], cattle [89], pigs [91], and musk shrews [92]. Thus, the POA/PeN kisspeptin neurons in those species are likely equivalent to AVPV-PeN kisspeptin neurons in rodents in terms of an estrogen positive feedback action site.
The notion that the AVPV-PeN kisspeptin neurons serve as an intrinsic source of the GnRH surge generator is more verified by the following studies on sex difference in LH surge generation in rodent models [129,130,131,132]. It is well-known that male rats failed to show LH surge even when they were treated with a preovulatory level of estrogen after castration in adulthood [133,134]. Concordantly, male rodents show only a few kisspeptin neurons in the AVPV-PeN even in the presence of estrogen, whereas females exhibit a cluster of AVPV-PeN kisspeptin neurons in the presence of estrogen [51,76,77]. Sex steroids originated from the perinatal testes are considered to cause defeminization of the AVPV-PeN kisspeptin neurons because neonatal castration allowed the estrogen-induced AVPV-PeN Kiss1 expression and LH surge in genetic male rats in adulthood to be shown [133,134]. In further support, neonatally androgenized/estrogenized female rats displayed the male-like pattern (few) of Kiss1 expression in the AVPV-PeN and failed to show LH surge in adulthood [76,133]. Thus, these findings suggest that the AVPV-PeN kisspeptin neurons serve as a target of the estrogen positive feedback action and are the intrinsic source of the GnRH surge generator in rodents.
Estrogen-induced Kiss1 expression in AVPV-PeN kisspeptin neurons is likely mediated via the ERE-dependent pathway because estrogen treatment failed to induce AVPV-PeN Kiss1 expression and LH surge generation in ERα KIKO mice [31,95]. In addition, our previous study using ChIP assay for ERα and acetylated histone H3 suggested that estrogen-bound ERα binds to the Kiss1 promoter and enhances histone H3 acetylation of the Kiss1 promoter region in the AVPV-PeN kisspeptin neurons because estrogen induces ERα binding and histone H3 acetylation at the Kiss1 promoter region in mouse AVPV-PeN tissue [98]. The finding suggests that an estrogen-dependent activating modification of histone H3 of the Kiss1 promoter resulted in an induction of Kiss1 expression. Furthermore, chromatin conformation capture (3C) assay suggested that estrogen induces chromatin loop formation between the Kiss1 promoter and the 3′-intergenic regions of the Kiss1 locus in AVPV-PeN kisspeptin neurons in mice [98]. This result also suggests that the 3′-intergenic region of the Kiss1 locus serves as an enhancer for estrogen-induced Kiss1 expression in AVPV-PeN kisspeptin neurons. Indeed, our in vivo reporter assay utilizing Kiss1-GFP reporter mice suggested that the 3′-intergenic region of the Kiss1 gene is required for an induction of Kiss1 mRNA expression by estrogen in the AVPV-PeN of female mice [98]. More specifically, the reporter mice carrying the 3′-truncated Kiss1-GFP transgene (RBRC09417) failed to display estrogen-induced GFP expression in AVPV-PeN kisspeptin neurons, but displayed ovariectomy-induced GFP expression in ARC kisspeptin neurons. As described above, the reporter mice carrying the full-length of Kiss1-GFP transgene (RBRC09413) display GFP expression in the AVPV-PeN kisspeptin neurons of the OVX mice with estrogen treatment and the ARC kisspeptin neurons without estrogen treatment.
We envisage here the molecular mechanism responsible for the estrogen positive feedback action on AVPV-PeN Kiss1 expression in rodents, as shown in Figure 3. Briefly, at proestrus in rodents, circulating high levels of estrogen bind to ERα in the AVPV-PeN kisspeptin neurons, and the estrogen-bound ERα may bind to ERE in the Kiss1 promoter region and enhance histone H3 acetylation at the promoter region. The estrogen-ERα bindings and/or the activating histone modification may form chromatin loops between the Kiss1 promoter and the 3′-intergenic regions of Kiss1 locus, resulting in Kiss1 expression in AVPV-PeN kisspeptin neurons.
Interestingly, both the proestrous and diestrous levels of estrogen are capable of increasing AVPV-PeN Kiss1 expression in female rats [51], while only the proestrous level of estrogen evoked LH surge in female rats [51]. This fact raises the possibility that, in addition to an increase in Kiss1 expression in AVPV-PeN kisspeptin neurons, certain afferent ERα-expressing neurons may be also involved in the positive feedback action of estrogen on kisspeptin release from the AVPV-PeN kisspeptin neurons. One of the candidates is brainstem noradrenergic neurons: previous studies showed that ERα expression was found in A2 noradrenergic neurons [135], where estrogen induced c-Fos expression [136], and that α1-adrenergic receptor antagonist attenuated afternoon LH surge in proestrous female rats [137]. Additionally, the SCN, where ERα mRNA expression was found in rats [33], might be also an estrogen positive feedback action site. It is well known that LH surge is timed by the circadian clock localized in the SCN and occurs in the afternoon of proestrus in rodents. A previous study suggested an involvement of SCN vasopressin neurons in the induction of afternoon LH surge because an administration of vasopressin V1 receptor antagonist attenuated afternoon LH surge in proestrous female rats [138]. Interestingly, an electrophysiological study showed that vasopressin treatment induced AVPV-PeN kisspeptin neuronal activity in estrogen-treated OVX mice but not in OVX mice [139], indicating that estrogen may enhance the sensitivity of AVPV-PeN kisspeptin neurons to vasopressin. Taken together, these findings suggest that A2 noradrenergic neurons and SCN vasopressin neurons may mediate estrogen positive feedback to induce kisspeptin release from AVPV-PeN kisspeptin neurons. Previous studies demonstrated that AVPV-PeN kisspeptin neurons largely project their axons to the GnRH cell bodies in the POA in mice [77,140] and that kisspeptin reportedly exerted a long-lasting excitation of GnRH neurons [141,142]. These findings suggest that kisspeptin secreted from the AVPV-PeN kisspeptin neurons may act on GnRH cell bodies to induce GnRH/LH surge.

6. Conclusions and Perspectives

Overall, the intensive studies on hypothalamic kisspeptin neurons in the past two decades have been gradually uncovering the cellular and molecular mechanisms of the negative and positive feedback actions of estrogen on GnRH pulse and surge generation in female mammals. Based on the findings currently available, we now postulate that the negative feedback action of estrogen, which fine-tunes GnRH pulses, is mainly mediated by the ARC kisspeptin neurons, in which estrogen directly represses Kiss1 expression. Further, estrogen may indirectly inhibit pulsatile kisspeptin release via likely afferent ERα-expressing neurons. Further studies are warranted to clarify the afferent inputs that convey estrogen signals to ARC kisspeptin neurons. In addition, we postulate that the positive feedback action of estrogen, which induces GnRH surge, is mainly mediated by the anterior population (the AVPV-PeN in rodents and the POA or PeN in other mammals) of hypothalamic kisspeptin neurons, in which estrogen directly induces Kiss1 expression. Furthermore, estrogen may indirectly stimulate surge-mode kisspeptin release via likely afferent ERα-expressing neurons, such as brainstem noradrenergic neurons and SCN vasopressin neurons. So far, only a few studies are available to show kisspeptin release, except for the studies in the rhesus monkeys [113,114], as described above. Further studies are needed to depict pulsatile and surge-modes of kisspeptin release and to clarify the mechanism of kisspeptin release controlled by the negative and positive feedback actions of estrogen.

Author Contributions

Writing—original draft preparation, Y.U. and H.T.; review and editing, N.I., S.N., and H.T.; visualization, S.N. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported in part by JSPS KAKENHI Grant Numbers, 20H03127 (to Y.U.); 19H03103 (to N.I.); 20K15650 (to S.N.); 18H03973, 21H05031, and 21K19186 (to H.T.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We respectfully acknowledge the contributions of the late Kei-ichiro Maeda, D.V.M., of The University of Tokyo, who suddenly passed away in February 2018. His leadership, insight, and invaluable ideas contributed greatly to our original studies reviewed in this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Harris, G.W.; Jacobsohn, D. Functional grafts of the anterior pituitary gland. Proc. R. Soc. Lond. B Biol. Sci. 1952, 139, 263–276. [Google Scholar] [PubMed]
  2. McCann, S.M. A hypothalamic luteinizing-hormone-releasing factor. Am. J. Physiol. 1962, 202, 395–400. [Google Scholar] [CrossRef]
  3. Ramirez, V.D.; McCann, S.M. A highly sensitive test for LH-releasing activity: The ovariectomized, estrogen progesterone-blocked rat. Endocrinology 1963, 73, 193–198. [Google Scholar] [CrossRef] [PubMed]
  4. Neill, J.D.; Johansson, E.D.; Datta, J.K.; Knobil, E. Relationship between the plasma levels of luteinizing hormone and progesterone during the normal menstrual cycle. J. Clin. Endocrinol. Metab. 1967, 27, 1167–1173. [Google Scholar] [CrossRef] [PubMed]
  5. Monroe, S.E.; Atkinson, L.E.; Knobil, E. Patterns of circulating luteinizing hormone and their relation to plasma progesterone levels during the menstrual cycle of the Rhesus monkey. Endocrinology 1970, 87, 453–455. [Google Scholar] [CrossRef]
  6. Atkinson, L.E.; Bhattacharya, A.N.; Monroe, S.E.; Dierschke, D.J.; Knobil, E. Effects of gonadectomy on plasma LH concentration in the rhesus monkey. Endocrinology 1970, 87, 847–849. [Google Scholar] [CrossRef]
  7. Yamaji, T.; Dierschke, D.J.; Hotchkiss, J.; Bhattacharya, A.N.; Surve, A.H.; Knobil, E. Estrogen induction of LH release in the rhesus monkey. Endocrinology 1971, 89, 1034–1041. [Google Scholar] [CrossRef]
  8. Karsch, F.J.; Dierschke, D.K.; Weick, R.F.; Yamaji, T.; Hotchkiss, J.; Knobil, E. Positive and negative feedback control by estrogen of luteinizing hormone secretion in the rhesus monkey. Endocrinology 1973, 92, 799–804. [Google Scholar] [CrossRef]
  9. Matsuo, H.; Baba, Y.; Nair, R.M.; Arimura, A.; Schally, A.V. Structure of the porcine LH- and FSH-releasing hormone. I. The proposed amino acid sequence. Biochem. Biophys. Res. Commun. 1971, 43, 1334–1339. [Google Scholar] [CrossRef]
  10. Amoss, M.; Burgus, R.; Blackwell, R.; Vale, W.; Fellows, R.; Guillemin, R. Purification, amino acid composition and N-terminus of the hypothalamic luteinizing hormone releasing factor (LRF) of ovine origin. Biochem. Biophys. Res. Commun. 1971, 44, 205–210. [Google Scholar] [CrossRef]
  11. Moenter, S.M.; Caraty, A.; Karsch, F.J. The estradiol-induced surge of gonadotropin-releasing hormone in the ewe. Endocrinology 1990, 127, 1375–1384. [Google Scholar] [CrossRef]
  12. Moenter, S.M.; Brand, R.M.; Midgley, A.R.; Karsch, F.J. Dynamics of gonadotropin-releasing hormone release during a pulse. Endocrinology 1992, 130, 503–510. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Moenter, S.M.; Brand, R.C.; Karsch, F.J. Dynamics of gonadotropin-releasing hormone (GnRH) secretion during the GnRH surge: Insights into the mechanism of GnRH surge induction. Endocrinology 1992, 130, 2978–2984. [Google Scholar] [CrossRef]
  14. Pau, K.Y.; Berria, M.; Hess, D.L.; Spies, H.G. Preovulatory gonadotropin-releasing hormone surge in ovarian-intact rhesus macaques. Endocrinology 1993, 133, 1650–1656. [Google Scholar] [CrossRef] [PubMed]
  15. Lincoln, D.W.; Fraser, H.M.; Lincoln, G.A.; Martin, G.B.; McNeilly, A.S. Hypothalamic pulse generators. Recent Prog. Horm. Res. 1985, 41, 369–419. [Google Scholar] [PubMed]
  16. Maeda, K.-I.; Tsukamura, H.; Ohkura, S.; Kawakami, S.; Nagabukuro, H.; Yokoyama, A. The LHRH pulse generator: A mediobasal hypothalamic location. Neurosci. Biobehav. Rev. 1995, 19, 427–437. [Google Scholar] [CrossRef]
  17. Kimura, F.; Funabashi, T. Two subgroups of gonadotropin-releasing hormone neurons control gonadotropin secretion in rats. News Physiol. Sci. 1998, 13, 225–231. [Google Scholar] [CrossRef] [PubMed]
  18. Couse, J.F.; Korach, K.S. Estrogen receptor null mice: What have we learned and where will they lead us? Endocr. Rev. 1999, 20, 358–417. [Google Scholar] [CrossRef] [PubMed]
  19. Couse, J.F.; Yates, M.M.; Walker, V.R.; Korach, K.S. Characterization of the hypothalamic-pituitary-gonadal axis in estrogen receptor (ER) Null mice reveals hypergonadism and endocrine sex reversal in females lacking ERα but not ERβ. Mol. Endocrinol. 2003, 17, 1039–1053. [Google Scholar] [CrossRef]
  20. Rumi, M.A.; Dhakal, P.; Kubota, K.; Chakraborty, D.; Lei, T.; Larson, M.A.; Wolfe, M.W.; Roby, K.F.; Vivian, J.L.; Soares, M.J. Generation of Esr1-knockout rats using zinc finger nuclease-mediated genome editing. Endocrinology 2014, 155, 1991–1999. [Google Scholar] [CrossRef] [Green Version]
  21. Antonson, P.; Apolinario, L.M.; Shamekh, M.M.; Humire, P.; Poutanen, M.; Ohlsson, C.; Nalvarte, I.; Gustafsson, J.A. Generation of an all-exon Esr2 deleted mouse line: Effects on fertility. Biochem. Biophys. Res. Commun. 2020, 529, 231–237. [Google Scholar] [CrossRef] [PubMed]
  22. Krege, J.H.; Hodgin, J.B.; Couse, J.F.; Enmark, E.; Warner, M.; Mahler, J.F.; Sar, M.; Korach, K.S.; Gustafsson, J.A.; Smithies, O. Generation and reproductive phenotypes of mice lacking estrogen receptor β. Proc. Natl. Acad. Sci. USA 1998, 95, 15677–15682. [Google Scholar] [CrossRef] [Green Version]
  23. Rumi, M.A.; Singh, P.; Roby, K.F.; Zhao, X.; Iqbal, K.; Ratri, A.; Lei, T.; Cui, W.; Borosha, S.; Dhakal, P.; et al. Defining the Role of Estrogen Receptor β in the Regulation of Female Fertility. Endocrinology 2017, 158, 2330–2343. [Google Scholar] [CrossRef] [Green Version]
  24. Roa, J.; Vigo, E.; Castellano, J.M.; Gaytan, F.; Navarro, V.M.; Aguilar, E.; Dijcks, F.A.; Ederveen, A.G.; Pinilla, L.; van Noort, P.I.; et al. Opposite roles of estrogen receptor (ER)-α and ERβ in the modulation of luteinizing hormone responses to kisspeptin in the female rat: Implications for the generation of the preovulatory surge. Endocrinology 2008, 149, 1627–1637. [Google Scholar] [CrossRef]
  25. Klinge, C.M. Estrogen receptor interaction with co-activators and co-repressors. Steroids 2000, 65, 227–251. [Google Scholar] [CrossRef]
  26. Stein, B.; Yang, M.X. Repression of the interleukin-6 promoter by estrogen receptor is mediated by NF-κB and C/EBPβ. Mol. Cell. Biol. 1995, 15, 4971–4979. [Google Scholar] [CrossRef] [Green Version]
  27. Ray, A.; Prefontaine, K.E.; Ray, P. Down-modulation of interleukin-6 gene expression by 17β-estradiol in the absence of high affinity DNA binding by the estrogen receptor. J. Biol. Chem. 1994, 269, 12940–12946. [Google Scholar] [CrossRef]
  28. Schmitt, M.; Bausero, P.; Simoni, P.; Queuche, D.; Geoffroy, V.; Marschal, C.; Kempf, J.; Quirin-Stricker, C. Positive and negative effects of nuclear receptors on transcription activation by AP-1 of the human choline acetyltransferase proximal promoter. J. Neurosci. Res. 1995, 40, 152–164. [Google Scholar] [CrossRef]
  29. Jakacka, M.; Ito, M.; Weiss, J.; Chien, P.Y.; Gehm, B.D.; Jameson, J.L. Estrogen receptor binding to DNA is not required for its activity through the nonclassical AP1 pathway. J. Biol. Chem. 2001, 276, 13615–13621. [Google Scholar] [CrossRef] [Green Version]
  30. Jakacka, M.; Ito, M.; Martinson, F.; Ishikawa, T.; Lee, E.J.; Jameson, J.L. An estrogen receptor (ER)α deoxyribonucleic acid-binding domain knock-in mutation provides evidence for nonclassical ER pathway signaling In Vivo. Mol. Endocrinol. 2002, 16, 2188–2201. [Google Scholar] [CrossRef] [Green Version]
  31. Glidewell-Kenney, C.; Hurley, L.A.; Pfaff, L.; Weiss, J.; Levine, J.E.; Jameson, J.L. Nonclassical estrogen receptor α signaling mediates negative feedback in the female mouse reproductive axis. Proc. Natl. Acad. Sci. USA 2007, 104, 8173–8177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Herbison, A.E.; Pape, J.R. New evidence for estrogen receptors in gonadotropin-releasing hormone neurons. Front. Neuroendocrinol. 2001, 22, 292–308. [Google Scholar] [CrossRef] [PubMed]
  33. Simerly, R.B.; Chang, C.; Muramatsu, M.; Swanson, L.W. Distribution of androgen and estrogen receptor mRNA-containing cells in the rat brain: An in situ hybridization study. J. Comp. Neurol. 1990, 294, 76–95. [Google Scholar] [CrossRef] [PubMed]
  34. Yokosuka, M.; Okamura, H.; Hayashi, S. Postnatal development and sex difference in neurons containing estrogen receptor-α immunoreactivity in the preoptic brain, the diencephalon, and the amygdala in the rat. J. Comp. Neurol. 1997, 389, 81–93. [Google Scholar] [CrossRef]
  35. Scott, C.J.; Tilbrook, A.J.; Simmons, D.M.; Rawson, J.A.; Chu, S.; Fuller, P.J.; Ing, N.H.; Clarke, I.J. The distribution of cells containing estrogen receptor-α (ERα) and ERβ messenger ribonucleic acid in the preoptic area and hypothalamus of the sheep: Comparison of males and females. Endocrinology 2000, 141, 2951–2962. [Google Scholar] [CrossRef]
  36. Lehman, M.N.; Ebling, F.J.; Moenter, S.M.; Karsch, F.J. Distribution of estrogen receptor-immunoreactive cells in the sheep brain. Endocrinology 1993, 133, 876–886. [Google Scholar] [CrossRef]
  37. Anderson, C.H.; Greenwald, G.S. Autoradiographic analysis of estradiol uptake in the brain and pituitary of the female rat. Endocrinology 1969, 85, 1160–1165. [Google Scholar] [CrossRef]
  38. Smith, E.R.; Davidson, J.M. Location of feedback receptors: Effects of intracranially implanted steroids on plasma LH and LRF response. Endocrinology 1974, 95, 1566–1573. [Google Scholar] [CrossRef]
  39. Akema, T.; Tadokoro, Y.; Kawakami, M. Changes in the characteristics of pulsatile LH secretion after estradiol implantation into the preoptic area and the basal hypothalamus in ovariectomized rats. Endocrinol. Jpn. 1983, 30, 281–287. [Google Scholar] [CrossRef] [Green Version]
  40. Nagatani, S.; Tsukamura, H.; Maeda, K.-I. Estrogen feedback needed at the paraventricular nucleus or A2 to suppress pulsatile luteinizing hormone release in fasting female rats. Endocrinology 1994, 135, 870–875. [Google Scholar] [CrossRef]
  41. Estacio, M.A.; Yamada, S.; Tsukamura, H.; Hirunagi, K.; Maeda, K. Effect of fasting and immobilization stress on estrogen receptor immunoreactivity in the brain in ovariectomized female rats. Brain Res. 1996, 717, 55–61. [Google Scholar] [CrossRef]
  42. Kawakami, M.; Konda, N.; Yoshioka, E. Stimulatory feedback action of estradiol and estrone on LH release in the ovariectomized rat: Roles different between limbic system and preoptic area. Endocrinol. Jpn. 1977, 24, 163–172. [Google Scholar] [CrossRef] [Green Version]
  43. Goodman, R.L. The site of the positive feedback action of estradiol in the rat. Endocrinology 1978, 102, 151–159. [Google Scholar] [CrossRef]
  44. Wiegand, S.J.; Terasawa, E.; Bridson, W.E.; Goy, R.W. Effects of discrete lesions of preoptic and suprachiasmatic structures in the female rat. Alterations in the feedback regulation of gonadotropin secretion. Neuroendocrinology 1980, 31, 147–157. [Google Scholar] [CrossRef]
  45. Wiegand, S.J.; Terasawa, E. Discrete lesions reveal functional heterogeneity of suprachiasmatic structures in regulation of gonadotropin secretion in the female rat. Neuroendocrinology 1982, 34, 395–404. [Google Scholar] [CrossRef]
  46. Petersen, S.L.; Cheuk, C.; Hartman, R.D.; Barraclough, C.A. Medial preoptic microimplants of the antiestrogen, keoxifene, affect luteinizing hormone-releasing hormone mRNA levels, median eminence luteinizing hormone-releasing hormone concentrations and luteinizing hormone release in ovariectomized, estrogen-treated rats. J. Neuroendocrinol. 1989, 1, 279–283. [Google Scholar]
  47. Petersen, S.L.; Barraclough, C.A. Suppression of spontaneous LH surges in estrogen-treated ovariectomized rats by microimplants of antiestrogens into the preoptic brain. Brain Res. 1989, 484, 279–289. [Google Scholar] [CrossRef]
  48. Smith, J.T.; Cunningham, M.J.; Rissman, E.F.; Clifton, D.K.; Steiner, R.A. Regulation of Kiss1 gene expression in the brain of the female mouse. Endocrinology 2005, 146, 3686–3692. [Google Scholar] [CrossRef] [PubMed]
  49. Smith, J.T.; Popa, S.M.; Clifton, D.K.; Hoffman, G.E.; Steiner, R.A. Kiss1 neurons in the forebrain as central processors for generating the preovulatory luteinizing hormone surge. J. Neurosci. 2006, 26, 6687–6694. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Kinoshita, M.; Tsukamura, H.; Adachi, S.; Matsui, H.; Uenoyama, Y.; Iwata, K.; Yamada, S.; Inoue, K.; Ohtaki, T.; Matsumoto, H.; et al. Involvement of central metastin in the regulation of preovulatory luteinizing hormone surge and estrous cyclicity in female rats. Endocrinology 2005, 146, 4431–4436. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Adachi, S.; Yamada, S.; Takatsu, Y.; Matsui, H.; Kinoshita, M.; Takase, K.; Sugiura, H.; Ohtaki, T.; Matsumoto, H.; Uenoyama, Y.; et al. Involvement of anteroventral periventricular metastin/kisspeptin neurons in estrogen positive feedback action on luteinizing hormone release in female rats. J. Reprod. Dev. 2007, 53, 367–378. [Google Scholar] [CrossRef] [Green Version]
  52. Franceschini, I.; Lomet, D.; Cateau, M.; Delsol, G.; Tillet, Y.; Caraty, A. Kisspeptin immunoreactive cells of the ovine preoptic area and arcuate nucleus co-express estrogen receptor alpha. Neurosci. Lett. 2006, 401, 225–230. [Google Scholar] [CrossRef] [PubMed]
  53. Gottsch, M.L.; Cunningham, M.J.; Smith, J.T.; Popa, S.M.; Acohido, B.V.; Crowley, W.F.; Seminara, S.; Clifton, D.K.; Steiner, R.A. A role for kisspeptins in the regulation of gonadotropin secretion in the mouse. Endocrinology 2004, 145, 4073–4077. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Irwig, M.S.; Fraley, G.S.; Smith, J.T.; Acohido, B.V.; Popa, S.M.; Cunningham, M.J.; Gottsch, M.L.; Clifton, D.K.; Steiner, R.A. Kisspeptin activation of gonadotropin releasing hormone neurons and regulation of KiSS-1 mRNA in the male rat. Neuroendocrinology 2004, 80, 264–272. [Google Scholar] [CrossRef] [PubMed]
  55. Matsui, H.; Takatsu, Y.; Kumano, S.; Matsumoto, H.; Ohtaki, T. Peripheral administration of metastin induces marked gonadotropin release and ovulation in the rat. Biochem. Biophys. Res. Commun. 2004, 320, 383–388. [Google Scholar] [CrossRef]
  56. Pheng, V.; Uenoyama, Y.; Homma, T.; Inamoto, Y.; Takase, K.; Yoshizawa-Kumagaye, K.; Isaka, S.; Watanabe, T.X.; Ohkura, S.; Tomikawa, J.; et al. Potencies of centrally- or peripherally-injected full-length kisspeptin or its C-terminal decapeptide on LH release in intact male rats. J. Reprod. Dev. 2009, 55, 378–382. [Google Scholar] [CrossRef] [Green Version]
  57. Navarro, V.M.; Castellano, J.M.; Fernandez-Fernandez, R.; Tovar, S.; Roa, J.; Mayen, A.; Nogueiras, R.; Vazquez, M.J.; Barreiro, M.L.; Magni, P.; et al. Characterization of the potent luteinizing hormone-releasing activity of KiSS-1 peptide, the natural ligand of GPR54. Endocrinology 2005, 146, 156–163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Navarro, V.M.; Castellano, J.M.; Fernandez-Fernandez, R.; Tovar, S.; Roa, J.; Mayen, A.; Barreiro, M.L.; Casanueva, F.F.; Aguilar, E.; Dieguez, C.; et al. Effects of KiSS-1 peptide, the natural ligand of GPR54, on follicle-stimulating hormone secretion in the rat. Endocrinology 2005, 146, 1689–1697. [Google Scholar] [CrossRef]
  59. Ohkura, S.; Takase, K.; Matsuyama, S.; Mogi, K.; Ichimaru, T.; Wakabayashi, Y.; Uenoyama, Y.; Mori, Y.; Steiner, R.A.; Tsukamura, H.; et al. Gonadotrophin-releasing hormone pulse generator activity in the hypothalamus of the goat. J. Neuroendocrinol. 2009, 21, 813–821. [Google Scholar] [CrossRef]
  60. Naniwa, Y.; Nakatsukasa, K.; Setsuda, S.; Oishi, S.; Fujii, N.; Matsuda, F.; Uenoyama, Y.; Tsukamura, H.; Maeda, K.-I.; Ohkura, S. Effects of full-length kisspeptin administration on follicular development in Japanese Black beef cows. J. Reprod. Dev. 2013, 59, 588–594. [Google Scholar] [CrossRef] [Green Version]
  61. Dhillo, W.S.; Chaudhri, O.B.; Patterson, M.; Thompson, E.L.; Murphy, K.G.; Badman, M.K.; McGowan, B.M.; Amber, V.; Patel, S.; Ghatei, M.A.; et al. Kisspeptin-54 stimulates the hypothalamic-pituitary gonadal axis in human males. J. Clin. Endocrinol. Metab. 2005, 90, 6609–6615. [Google Scholar] [CrossRef] [Green Version]
  62. Shahab, M.; Mastronardi, C.; Seminara, S.B.; Crowley, W.F.; Ojeda, S.R.; Plant, T.M. Increased hypothalamic GPR54 signaling: A potential mechanism for initiation of puberty in primates. Proc. Natl. Acad. Sci. USA 2005, 102, 2129–2134. [Google Scholar] [CrossRef] [Green Version]
  63. Ohtaki, T.; Shintani, Y.; Honda, S.; Matsumoto, H.; Hori, A.; Kanehashi, K.; Terao, Y.; Kumano, S.; Takatsu, Y.; Masuda, Y.; et al. Metastasis suppressor gene KiSS-1 encodes peptide ligand of a G-protein-coupled receptor. Nature 2001, 411, 613–617. [Google Scholar] [CrossRef]
  64. Kotani, M.; Detheux, M.; Vandenbogaerde, A.; Communi, D.; Vanderwinden, J.M.; Le Poul, E.; Brezillon, S.; Tyldesley, R.; Suarez-Huerta, N.; Vandeput, F.; et al. The metastasis suppressor gene KiSS-1 encodes kisspeptins, the natural ligands of the orphan G protein-coupled receptor GPR54. J. Biol. Chem. 2001, 276, 34631–34636. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. de Roux, N.; Genin, E.; Carel, J.C.; Matsuda, F.; Chaussain, J.L.; Milgrom, E. Hypogonadotropic hypogonadism due to loss of function of the KiSS1-derived peptide receptor GPR54. Proc. Natl. Acad. Sci. USA 2003, 100, 10972–10976. [Google Scholar] [CrossRef] [Green Version]
  66. Seminara, S.B.; Messager, S.; Chatzidaki, E.E.; Thresher, R.R.; Acierno, J.S., Jr.; Shagoury, J.K.; Bo-Abbas, Y.; Kuohung, W.; Schwinof, K.M.; Hendrick, A.G.; et al. The GPR54 gene as a regulator of puberty. N. Engl. J. Med. 2003, 349, 1614–1627. [Google Scholar] [CrossRef] [Green Version]
  67. Topaloglu, A.K.; Tello, J.A.; Kotan, L.D.; Ozbek, M.N.; Yilmaz, M.B.; Erdogan, S.; Gurbuz, F.; Temiz, F.; Millar, R.P.; Yuksel, B. Inactivating KISS1 mutation and hypogonadotropic hypogonadism. N. Engl. J. Med. 2012, 366, 629–635. [Google Scholar] [CrossRef] [Green Version]
  68. Messager, S.; Chatzidaki, E.E.; Ma, D.; Hendrick, A.G.; Zahn, D.; Dixon, J.; Thresher, R.R.; Malinge, I.; Lomet, D.; Carlton, M.B.; et al. Kisspeptin directly stimulates gonadotropin-releasing hormone release via G protein-coupled receptor 54. Proc. Natl. Acad. Sci. USA 2005, 102, 1761–1766. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. d’Anglemont de Tassigny, X.; Fagg, L.A.; Dixon, J.P.; Day, K.; Leitch, H.G.; Hendrick, A.G.; Zahn, D.; Franceschini, I.; Caraty, A.; Carlton, M.B.; et al. Hypogonadotropic hypogonadism in mice lacking a functional Kiss1 gene. Proc. Natl. Acad. Sci. USA 2007, 104, 10714–10719. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Lapatto, R.; Pallais, J.C.; Zhang, D.; Chan, Y.M.; Mahan, A.; Cerrato, F.; Le, W.W.; Hoffman, G.E.; Seminara, S.B. Kiss1-/- mice exhibit more variable hypogonadism than Gpr54-/- mice. Endocrinology 2007, 148, 4927–4936. [Google Scholar] [CrossRef]
  71. Chan, Y.M.; Broder-Fingert, S.; Wong, K.M.; Seminara, S.B. Kisspeptin/Gpr54-independent gonadotrophin-releasing hormone activity in Kiss1 and Gpr54 mutant mice. J. Neuroendocrinol. 2009, 21, 1015–1023. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Uenoyama, Y.; Nakamura, S.; Hayakawa, Y.; Ikegami, K.; Watanabe, Y.; Deura, C.; Minabe, S.; Tomikawa, J.; Goto, T.; Ieda, N.; et al. Lack of pulse and surge modes and glutamatergic stimulation of LH release in Kiss1 knockout rats. J. Neuroendocrinol. 2015, 27, 187–197. [Google Scholar] [CrossRef] [PubMed]
  73. Higo, S.; Honda, S.; Iijima, N.; Ozawa, H. Mapping of kisspeptin receptor mRNA in the whole rat brain and its co-localisation with oxytocin in the paraventricular nucleus. J. Neuroendocrinol. 2016, 28. [Google Scholar] [CrossRef] [PubMed]
  74. Herbison, A.E.; de Tassigny, X.; Doran, J.; Colledge, W.H. Distribution and postnatal development of Gpr54 gene expression in mouse brain and gonadotropin-releasing hormone neurons. Endocrinology 2010, 151, 312–321. [Google Scholar] [CrossRef] [Green Version]
  75. Han, S.K.; Gottsch, M.L.; Lee, K.J.; Popa, S.M.; Smith, J.T.; Jakawich, S.K.; Clifton, D.K.; Steiner, R.A.; Herbison, A.E. Activation of gonadotropin-releasing hormone neurons by kisspeptin as a neuroendocrine switch for the onset of puberty. J. Neurosci. 2005, 25, 11349–11356. [Google Scholar] [CrossRef]
  76. Kauffman, A.S.; Gottsch, M.L.; Roa, J.; Byquist, A.C.; Crown, A.; Clifton, D.K.; Hoffman, G.E.; Steiner, R.A.; Tena-Sempere, M. Sexual differentiation of Kiss1 gene expression in the brain of the rat. Endocrinology 2007, 148, 1774–1783. [Google Scholar] [CrossRef]
  77. Clarkson, J.; Herbison, A.E. Postnatal development of kisspeptin neurons in mouse hypothalamus; sexual dimorphism and projections to gonadotropin-releasing hormone neurons. Endocrinology 2006, 147, 5817–5825. [Google Scholar] [CrossRef]
  78. Takase, K.; Uenoyama, Y.; Inoue, N.; Matsui, H.; Yamada, S.; Shimizu, M.; Homma, T.; Tomikawa, J.; Kanda, S.; Matsumoto, H.; et al. Possible role of oestrogen in pubertal increase of Kiss1/kisspeptin expression in discrete hypothalamic areas of female rats. J. Neuroendocrinol. 2009, 21, 527–537. [Google Scholar] [CrossRef]
  79. Hrabovszky, E.; Takacs, S.; Gocz, B.; Skrapits, K. New Perspectives for Anatomical and Molecular Studies of Kisspeptin Neurons in the Aging Human Brain. Neuroendocrinology 2019, 109, 230–241. [Google Scholar] [CrossRef] [Green Version]
  80. Rometo, A.M.; Krajewski, S.J.; Voytko, M.L.; Rance, N.E. Hypertrophy and increased kisspeptin gene expression in the hypothalamic infundibular nucleus of postmenopausal women and ovariectomized monkeys. J. Clin. Endocrinol. Metab. 2007, 92, 2744–2750. [Google Scholar] [CrossRef] [Green Version]
  81. Smith, J.T.; Shahab, M.; Pereira, A.; Pau, K.Y.; Clarke, I.J. Hypothalamic expression of KISS1 and gonadotropin inhibitory hormone genes during the menstrual cycle of a non-human primate. Biol. Reprod. 2010, 83, 568–577. [Google Scholar] [CrossRef]
  82. Ramaswamy, S.; Guerriero, K.A.; Gibbs, R.B.; Plant, T.M. Structural interactions between kisspeptin and GnRH neurons in the mediobasal hypothalamus of the male rhesus monkey (Macaca mulatta) as revealed by double immunofluorescence and confocal microscopy. Endocrinology 2008, 149, 4387–4395. [Google Scholar] [CrossRef] [PubMed]
  83. Watanabe, Y.; Uenoyama, Y.; Suzuki, J.; Takase, K.; Suetomi, Y.; Ohkura, S.; Inoue, N.; Maeda, K.-I.; Tsukamura, H. Oestrogen-induced activation of preoptic kisspeptin neurones may be involved in the luteinizing hormone surge in male and female Japanese monkeys. J. Neuroendocrinol. 2014, 26, 909–917. [Google Scholar] [CrossRef] [PubMed]
  84. Estrada, K.M.; Clay, C.M.; Pompolo, S.; Smith, J.T.; Clarke, I.J. Elevated KiSS-1 expression in the arcuate nucleus prior to the cyclic preovulatory gonadotrophin-releasing hormone/lutenising hormone surge in the ewe suggests a stimulatory role for kisspeptin in oestrogen-positive feedback. J. Neuroendocrinol. 2006, 18, 806–809. [Google Scholar] [CrossRef]
  85. Smith, J.T.; Li, Q.; Pereira, A.; Clarke, I.J. Kisspeptin neurons in the ovine arcuate nucleus and preoptic area are involved in the preovulatory luteinizing hormone surge. Endocrinology 2009, 150, 5530–5538. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Goodman, R.L.; Lehman, M.N.; Smith, J.T.; Coolen, L.M.; de Oliveira, C.V.; Jafarzadehshirazi, M.R.; Pereira, A.; Iqbal, J.; Caraty, A.; Ciofi, P.; et al. Kisspeptin neurons in the arcuate nucleus of the ewe express both dynorphin A and neurokinin B. Endocrinology 2007, 148, 5752–5760. [Google Scholar] [CrossRef]
  87. Wakabayashi, Y.; Nakada, T.; Murata, K.; Ohkura, S.; Mogi, K.; Navarro, V.M.; Clifton, D.K.; Mori, Y.; Tsukamura, H.; Maeda, K.-I.; et al. Neurokinin B and dynorphin A in kisspeptin neurons of the arcuate nucleus participate in generation of periodic oscillation of neural activity driving pulsatile gonadotropin-releasing hormone secretion in the goat. J. Neurosci. 2010, 30, 3124–3132. [Google Scholar] [CrossRef] [PubMed]
  88. Matsuda, F.; Nakatsukasa, K.; Suetomi, Y.; Naniwa, Y.; Ito, D.; Inoue, N.; Wakabayashi, Y.; Okamura, H.; Maeda, K.-I.; Uenoyama, Y.; et al. The LH surge-generating system is functional in male goats as in females: Involvement of kisspeptin neurones in the medial preoptic area. J. Neuroendocrinol. 2015, 27, 57–65. [Google Scholar] [CrossRef]
  89. Hassaneen, A.S.A.; Naniwa, Y.; Suetomi, Y.; Matsuyama, S.; Kimura, K.; Ieda, N.; Inoue, N.; Uenoyama, Y.; Tsukamura, H.; Maeda, K.-I.; et al. Immunohistochemical characterization of the arcuate kisspeptin/neurokinin B/dynorphin (KNDy) and preoptic kisspeptin neuronal populations in the hypothalamus during the estrous cycle in heifers. J. Reprod. Dev. 2016, 62, 471–477. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Decourt, C.; Tillet, Y.; Caraty, A.; Franceschini, I.; Briant, C. Kisspeptin immunoreactive neurons in the equine hypothalamus Interactions with GnRH neuronal system. J. Chem. Neuroanat. 2008, 36, 131–137. [Google Scholar] [CrossRef]
  91. Tomikawa, J.; Homma, T.; Tajima, S.; Shibata, T.; Inamoto, Y.; Takase, K.; Inoue, N.; Ohkura, S.; Uenoyama, Y.; Maeda, K.-I.; et al. Molecular characterization and estrogen regulation of hypothalamic KISS1 gene in the pig. Biol. Reprod. 2010, 82, 313–319. [Google Scholar] [CrossRef]
  92. Inoue, N.; Sasagawa, K.; Ikai, K.; Sasaki, Y.; Tomikawa, J.; Oishi, S.; Fujii, N.; Uenoyama, Y.; Ohmori, Y.; Yamamoto, N.; et al. Kisspeptin neurons mediate reflex ovulation in the musk shrew (Suncus murinus). Proc. Natl. Acad. Sci. USA 2011, 108, 17527–17532. [Google Scholar] [CrossRef] [Green Version]
  93. Smith, J.T.; Clay, C.M.; Caraty, A.; Clarke, I.J. KiSS-1 messenger ribonucleic acid expression in the hypothalamus of the ewe is regulated by sex steroids and season. Endocrinology 2007, 148, 1150–1157. [Google Scholar] [CrossRef]
  94. Smith, J.T.; Coolen, L.M.; Kriegsfeld, L.J.; Sari, I.P.; Jaafarzadehshirazi, M.R.; Maltby, M.; Bateman, K.; Goodman, R.L.; Tilbrook, A.J.; Ubuka, T.; et al. 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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Gottsch, M.L.; Navarro, V.M.; Zhao, Z.; Glidewell-Kenney, C.; Weiss, J.; Jameson, J.L.; Clifton, D.K.; Levine, J.E.; Steiner, R.A. Regulation of Kiss1 and Dynorphin gene expression in the murine brain by classical and nonclassical estrogen receptor pathways. J. Neurosci. 2009, 29, 9390–9395. [Google Scholar] [CrossRef]
  96. Yang, J.A.; Mamounis, K.J.; Yasrebi, A.; Roepke, T.A. Regulation of gene expression by 17β-estradiol in the arcuate nucleus of the mouse through ERE-dependent and ERE-independent mechanisms. Steroids 2016, 107, 128–138. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Yang, J.A.; Stires, H.; Belden, W.J.; Roepke, T.A. The arcuate estrogen-regulated transcriptome: Estrogen response element-dependent and -independent signaling of ERα in female mice. Endocrinology 2017, 158, 612–626. [Google Scholar] [PubMed]
  98. Tomikawa, J.; Uenoyama, Y.; Ozawa, M.; Fukanuma, T.; Takase, K.; Goto, T.; Abe, H.; Ieda, N.; Minabe, S.; Deura, C.; et al. Epigenetic regulation of Kiss1 gene expression mediating estrogen-positive feedback action in the mouse brain. Proc. Natl. Acad. Sci. USA 2012, 109, E1294–E1301. [Google Scholar] [CrossRef] [Green Version]
  99. Goto, T.; Tomikawa, J.; Ikegami, K.; Minabe, S.; Abe, H.; Fukanuma, T.; Imamura, T.; Takase, K.; Sanbo, M.; Tomita, K.; et al. Identification of hypothalamic arcuate nucleus-specific enhancer region of Kiss1 gene in mice. Mol. Endocrinol. 2015, 29, 121–129. [Google Scholar] [CrossRef] [Green Version]
  100. Lehman, M.N.; Coolen, L.M.; Goodman, R.L. Minireview: Kisspeptin/neurokinin B/dynorphin (KNDy) cells of the arcuate nucleus: A central node in the control of gonadotropin-releasing hormone secretion. Endocrinology 2010, 151, 3479–3489. [Google Scholar] [CrossRef]
  101. Murakawa, H.; Iwata, K.; Takeshita, T.; Ozawa, H. Immunoelectron microscopic observation of the subcellular localization of kisspeptin, neurokinin B and dynorphin A in KNDy neurons in the arcuate nucleus of the female rat. Neurosci. Lett. 2016, 612, 161–166. [Google Scholar] [CrossRef]
  102. Navarro, V.M.; Gottsch, M.L.; Chavkin, C.; Okamura, H.; Clifton, D.K.; Steiner, R.A. Regulation of gonadotropin-releasing hormone secretion by kisspeptin/dynorphin/neurokinin B neurons in the arcuate nucleus of the mouse. J. Neurosci. 2009, 29, 11859–11866. [Google Scholar] [CrossRef] [PubMed]
  103. Maeda, K.; Ohkura, S.; Uenoyama, Y.; Wakabayashi, Y.; Oka, Y.; Tsukamura, H.; Okamura, H. Neurobiological mechanisms underlying GnRH pulse generation by the hypothalamus. Brain Res. 2010, 1364, 103–115. [Google Scholar] [CrossRef] [PubMed]
  104. Okamura, H.; Tsukamura, H.; Ohkura, S.; Uenoyama, Y.; Wakabayashi, Y.; Maeda, K.-I. Kisspeptin and GnRH Pulse Generation. In Kisspeptin Signaling in Reproductive Biology; Kauffman, A.S., Smith, J.T., Eds.; Springer: New York, NY, USA, 2013; pp. 297–323. [Google Scholar]
  105. Goodman, R.L.; Ohkura, S.; Okamura, H.; Coolen, L.M.; Lehman, M.N. KNDy hypothesis for generation of GnRH pulses: Evidence from sheep and goats. In The GnRH Neuron and Its Control; Herbison, A.E., Plant, T.M., Eds.; Wiley: Hoboken, NJ, USA, 2018; pp. 289–324. [Google Scholar]
  106. Herbison, A.E. The gonadotropin-releasing hormone pulse generator. Endocrinology 2018, 159, 3723–3736. [Google Scholar] [CrossRef] [Green Version]
  107. Uenoyama, Y.; Pheng, V.; Tsukamura, H.; Maeda, K.I. The roles of kisspeptin revisited: Inside and outside the hypothalamus. J. Reprod. Dev. 2016, 62, 537–545. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Nagae, M.; Uenoyama, Y.; Okamoto, S.; Tsuchida, H.; Ikegami, K.; Goto, T.; Majarune, S.; Nakamura, S.; Sanbo, M.; Hirabayashi, M.; et al. Direct evidence that KNDy neurons maintain gonadotropin pulses and folliculogenesis as the GnRH pulse generator. Proc. Natl. Acad. Sci. USA 2021, 118, e2009156118. [Google Scholar] [CrossRef]
  109. Ikegami, K.; Goto, T.; Nakamura, S.; Watanabe, Y.; Sugimoto, A.; Majarune, S.; Horihata, K.; Nagae, M.; Tomikawa, J.; Imamura, T.; et al. Conditional kisspeptin neuron-specific Kiss1 knockout with newly generated Kiss1-floxed and Kiss1-Cre mice replicates a hypogonadal phenotype of global Kiss1 knockout mice. J. Reprod. Dev. 2020, 66, 359–367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Minabe, S.; Nakamura, S.; Fukushima, E.; Sato, M.; Ikegami, K.; Goto, T.; Sanbo, M.; Hirabayashi, M.; Tomikawa, J.; Imamura, T.; et al. Inducible Kiss1 knockdown in the hypothalamic arcuate nucleus suppressed pulsatile secretion of luteinizing hormone in male mice. J. Reprod. Dev. 2020, 66, 369–375. [Google Scholar] [CrossRef] [Green Version]
  111. Clarkson, J.; Han, S.Y.; Piet, R.; McLennan, T.; Kane, G.M.; Ng, J.; Porteous, R.W.; Kim, J.S.; Colledge, W.H.; Iremonger, K.J.; et al. Definition of the hypothalamic GnRH pulse generator in mice. Proc. Natl. Acad. Sci. USA 2017, 114, E10216–E10223. [Google Scholar] [CrossRef] [Green Version]
  112. Han, S.Y.; Kane, G.; Cheong, I.; Herbison, A.E. Characterization of GnRH Pulse Generator Activity in Male Mice Using GCaMP Fiber Photometry. Endocrinology 2019, 160, 557–567. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Keen, K.L.; Wegner, F.H.; Bloom, S.R.; Ghatei, M.A.; Terasawa, E. An increase in kisspeptin-54 release occurs with the pubertal increase in luteinizing hormone-releasing hormone-1 release in the stalk-median eminence of female rhesus monkeys In Vivo. Endocrinology 2008, 149, 4151–4157. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Kurian, J.R.; Keen, K.L.; Guerriero, K.A.; Terasawa, E. Tonic control of kisspeptin release in prepubertal monkeys: Implications to the mechanism of puberty onset. Endocrinology 2012, 153, 3331–3336. [Google Scholar] [CrossRef]
  115. Ikegami, K.; Minabe, S.; Ieda, N.; Goto, T.; Sugimoto, A.; Nakamura, S.; Inoue, N.; Oishi, S.; Maturana, A.D.; Sanbo, M.; et al. Evidence of involvement of neurone-glia/neurone-neurone communications via gap junctions in synchronised activity of KNDy neurones. J. Neuroendocrinol. 2017, 29. [Google Scholar] [CrossRef]
  116. Navarro, V.M.; Gottsch, M.L.; Wu, M.; Garcia-Galiano, D.; Hobbs, S.J.; Bosch, M.A.; Pinilla, L.; Clifton, D.K.; Dearth, A.; Ronnekleiv, O.K.; et al. Regulation of NKB pathways and their roles in the control of Kiss1 neurons in the arcuate nucleus of the male mouse. Endocrinology 2011, 152, 4265–4275. [Google Scholar] [CrossRef] [Green Version]
  117. Ruka, K.A.; Burger, L.L.; Moenter, S.M. Regulation of arcuate neurons coexpressing kisspeptin, neurokinin B, and dynorphin by modulators of neurokinin 3 and kappa-opioid receptors in adult male mice. Endocrinology 2013, 154, 2761–2771. [Google Scholar] [CrossRef] [Green Version]
  118. Tsuchida, H.; Mostari, P.; Yamada, K.; Miyazaki, S.; Enomoto, Y.; Inoue, N.; Uenoyama, Y.; Tsukamura, H. Paraventricular dynorphin A neurons mediate LH pulse suppression induced by hindbrain glucoprivation in female rats. Endocrinology 2020, 161, bqaa161. [Google Scholar] [CrossRef] [PubMed]
  119. Assadullah; Ieda, N.; Kawai, N.; Ishii, H.; Ihara, K.; Inoue, N.; Uenoyama, Y.; Tsukamura, H. Co-expression of the calcitonin receptor gene in the hypothalamic kisspeptin neurons in female rats. Reprod. Med. Biol. 2018, 17, 164–172. [Google Scholar] [CrossRef] [PubMed]
  120. Weems, P.W.; Witty, C.F.; Amstalden, M.; Coolen, L.M.; Goodman, R.L.; Lehman, M.N. Kappa opioid receptor is co-localized in GnRH and KNDy cells in the female ovine and rat brain. Endocrinology 2016, 157, 2367–2379. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Amstalden, M.; Coolen, L.M.; Hemmerle, A.M.; Billings, H.J.; Connors, J.M.; Goodman, R.L.; Lehman, M.N. Neurokinin 3 receptor immunoreactivity in the septal region, preoptic area and hypothalamus of the female sheep: Colocalisation in neurokinin B cells of the arcuate nucleus but not in gonadotrophin-releasing hormone neurones. J. Neuroendocrinol. 2010, 22, 1–12. [Google Scholar] [CrossRef] [Green Version]
  122. Dellovade, T.L.; Merchenthaler, I. Estrogen regulation of neurokinin B gene expression in the mouse arcuate nucleus is mediated by estrogen receptor α. Endocrinology 2004, 145, 736–742. [Google Scholar] [CrossRef] [Green Version]
  123. Pillon, D.; Caraty, A.; Fabre-Nys, C.; Bruneau, G. Short-term effect of oestradiol on neurokinin B mRNA expression in the infundibular nucleus of ewes. J. Neuroendocrinol. 2003, 15, 749–753. [Google Scholar] [CrossRef] [PubMed]
  124. Abel, T.W.; Voytko, M.L.; Rance, N.E. The effects of hormone replacement therapy on hypothalamic neuropeptide gene expression in a primate model of menopause. J. Clin. Endocrinol. Metab. 1999, 84, 2111–2118. [Google Scholar] [CrossRef]
  125. Kanaya, M.; Iwata, K.; Ozawa, H. Distinct dynorphin expression patterns with low- and high-dose estrogen treatment in the arcuate nucleus of female rats. Biol. Reprod. 2017, 97, 709–718. [Google Scholar] [CrossRef] [PubMed]
  126. Dubois, S.L.; Acosta-Martinez, M.; DeJoseph, M.R.; Wolfe, A.; Radovick, S.; Boehm, U.; Urban, J.H.; Levine, J.E. Positive, but not negative feedback actions of estradiol in adult female mice require estrogen receptor α in kisspeptin neurons. Endocrinology 2015, 156, 1111–1120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Mostari, M.P.; Ieda, N.; Deura, C.; Minabe, S.; Yamada, S.; Uenoyama, Y.; Maeda, K.-I.; Tsukamura, H. Dynorphin-kappa opioid receptor signaling partly mediates estrogen negative feedback effect on LH pulses in female rats. J. Reprod. Dev. 2013, 59, 266–272. [Google Scholar] [CrossRef] [Green Version]
  128. Stephens, S.B.Z.; Kauffman, A.S. Estrogen regulation of the molecular phenotype and active translatome of AVPV kisspeptin neurons. Endocrinology 2021, 162, bqab080. [Google Scholar] [CrossRef]
  129. Clarkson, J.; Herbison, A.E. Oestrogen, kisspeptin, GPR54 and the preovulatory luteinising hormone surge. J. Neuroendocrinol. 2009, 21, 305–311. [Google Scholar] [CrossRef]
  130. Tsukamura, H. Kobayashi Award 2019: The neuroendocrine regulation of the mammalian reproduction. Gen. Comp. Endocrinol. 2021, 113755. [Google Scholar] [CrossRef]
  131. Tsukamura, H.; Homma, T.; Tomikawa, J.; Uenoyama, Y.; Maeda, K.-I. Sexual differentiation of kisspeptin neurons responsible for sex difference in gonadotropin release in rats. Ann. N. Y. Acad. Sci. 2010, 1200, 95–103. [Google Scholar] [CrossRef]
  132. Tsukamura, H.; Maeda, K.-I.; Uenoyama, Y. Fetal/perinatal programming causing sexual dimorphism of the kisspeptin–GnRH neuronal network. In The GnRH Neuron and Its Control; Herbison, A.E., Plant, T.M., Eds.; Wiley: Hoboken, NJ, USA, 2018; pp. 43–60. [Google Scholar]
  133. Homma, T.; Sakakibara, M.; Yamada, S.; Kinoshita, M.; Iwata, K.; Tomikawa, J.; Kanazawa, T.; Matsui, H.; Takatsu, Y.; Ohtaki, T.; et al. Significance of neonatal testicular sex steroids to defeminize anteroventral periventricular kisspeptin neurons and the GnRH/LH surge system in male rats. Biol. Reprod. 2009, 81, 1216–1225. [Google Scholar] [CrossRef]
  134. Sakakibara, M.; Deura, C.; Minabe, S.; Iwata, Y.; Uenoyama, Y.; Maeda, K.I.; Tsukamura, H. Different critical perinatal periods and hypothalamic sites of oestradiol action in the defeminization of LH surge and lordosis capacity in the rat. J. Neuroendocrinol. 2013, 25, 251–259. [Google Scholar] [CrossRef]
  135. Simonian, S.X.; Herbison, A.E. Differential expression of estrogen receptor and neuropeptide Y by brainstem A1 and A2 noradrenaline neurons. Neuroscience 1997, 76, 517–529. [Google Scholar] [CrossRef]
  136. Jennes, L.; Jennes, M.E.; Purvis, C.; Nees, M. c-fos expression in noradrenergic A2 neurons of the rat during the estrous cycle and after steroid hormone treatments. Brain Res. 1992, 586, 171–175. [Google Scholar] [CrossRef]
  137. Le, W.W.; Berghorn, K.A.; Smith, M.S.; Hoffman, G.E. Alpha1-adrenergic receptor blockade blocks LH secretion but not LHRH cFos activation. Brain Res. 1997, 747, 236–245. [Google Scholar] [CrossRef]
  138. Funabashi, T.; Aiba, S.; Sano, A.; Shinohara, K.; Kimura, F. Intracerebroventricular injection of arginine-vasopressin V1 receptor antagonist attenuates the surge of luteinizing hormone and prolactin secretion in proestrous rats. Neurosci. Lett. 1999, 260, 37–40. [Google Scholar] [CrossRef]
  139. Piet, R.; Fraissenon, A.; Boehm, U.; Herbison, A.E. Estrogen permits vasopressin signaling in preoptic kisspeptin neurons in the female mouse. J. Neurosci. 2015, 35, 6881–6892. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Yip, S.H.; Boehm, U.; Herbison, A.E.; Campbell, R.E. Conditional viral tract tracing delineates the projections of the distinct kisspeptin neuron populations to gonadotropin-releasing hormone (GnRH) neurons in the mouse. Endocrinology 2015, 156, 2582–2594. [Google Scholar] [CrossRef] [Green Version]
  141. Zhang, C.; Roepke, T.A.; Kelly, M.J.; Ronnekleiv, O.K. Kisspeptin depolarizes gonadotropin-releasing hormone neurons through activation of TRPC-like cationic channels. J. Neurosci. 2008, 28, 4423–4434. [Google Scholar] [CrossRef]
  142. Liu, X.; Lee, K.; Herbison, A.E. Kisspeptin excites gonadotropin-releasing hormone neurons through a phospholipase C/calcium-dependent pathway regulating multiple ion channels. Endocrinology 2008, 149, 4605–4614. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Central mechanisms underlying the negative and positive feedback actions of estrogen on pulsatile and surge modes of gonadotropin-releasing hormone (GnRH)/luteinizing hormone (LH) release in female rodents. Estrogen production along with follicular development is stimulated by GnRH/gonadotropin pulses. During the follicular development period, low levels of circulating estrogen fine-tune GnRH/H pulses via the negative feedback action of estrogen. The estrogen negative feedback action is considered to be mediated by estrogen receptor α (ERα)-expressing kisspeptin neurons located in the arcuate nucleus (ARC). Estrogen production and release gradually increase along with the follicular development, and consequent high levels of circulating estrogen derived from mature follicles, in turn, induce GnRH/LH surge and hence ovulation via the positive feedback action of estrogen. The estrogen positive feedback action is likely mediated by ERα-expressing kisspeptin neurons located in the anteroventral periventricular nucleus–periventricular nucleus continuum (AVPV-PeN).
Figure 1. Central mechanisms underlying the negative and positive feedback actions of estrogen on pulsatile and surge modes of gonadotropin-releasing hormone (GnRH)/luteinizing hormone (LH) release in female rodents. Estrogen production along with follicular development is stimulated by GnRH/gonadotropin pulses. During the follicular development period, low levels of circulating estrogen fine-tune GnRH/H pulses via the negative feedback action of estrogen. The estrogen negative feedback action is considered to be mediated by estrogen receptor α (ERα)-expressing kisspeptin neurons located in the arcuate nucleus (ARC). Estrogen production and release gradually increase along with the follicular development, and consequent high levels of circulating estrogen derived from mature follicles, in turn, induce GnRH/LH surge and hence ovulation via the positive feedback action of estrogen. The estrogen positive feedback action is likely mediated by ERα-expressing kisspeptin neurons located in the anteroventral periventricular nucleus–periventricular nucleus continuum (AVPV-PeN).
Ijms 22 09229 g001
Figure 2. Putative molecular mechanism of the negative feedback action of estrogen on Kiss1 expression in the arcuate nucleus (ARC). Circulating estrogen seems to act on ARC kisspeptin neurons, in which estrogen-bound estrogen receptor α (ERα) coupled with an unknown transcriptional partner may repress Kiss1 expression via histone deacetylation and unwinding chromatin loops between the Kiss1 promoter and the 5′-intergenic regions of Kiss1 locus. In the absence of estrogen, ARC Kiss1 expression may be up-regulated by histone acetylation and chromatin loop formation between the Kiss1 promoter and the 5′-intergenic regions of the Kiss1 locus.
Figure 2. Putative molecular mechanism of the negative feedback action of estrogen on Kiss1 expression in the arcuate nucleus (ARC). Circulating estrogen seems to act on ARC kisspeptin neurons, in which estrogen-bound estrogen receptor α (ERα) coupled with an unknown transcriptional partner may repress Kiss1 expression via histone deacetylation and unwinding chromatin loops between the Kiss1 promoter and the 5′-intergenic regions of Kiss1 locus. In the absence of estrogen, ARC Kiss1 expression may be up-regulated by histone acetylation and chromatin loop formation between the Kiss1 promoter and the 5′-intergenic regions of the Kiss1 locus.
Ijms 22 09229 g002
Figure 3. Putative molecular mechanism of the estrogen positive feedback action on Kiss1 expression in the anteroventral-periventricular nucleus-periventricular nucleus continuum (AVPV-PeN). Preovulatory levels of circulating estrogen seem to act on AVPV-PeN kisspeptin neurons, in which estrogen-bound estrogen receptor α (ERα) may increase Kiss1 expression via histone acetylation of Kiss1 promotor region and chromatin loop formation between the Kiss1 promoter and the 3′-intergenic regions of Kiss1 locus. In the absence of estrogen, AVPV-PeN Kiss1 expression may be down-regulated by histone deacetylation of the Kiss1 promotor region and unwinding chromatin loops between the Kiss1 promoter and the 3′-intergenic regions of the Kiss1 locus.
Figure 3. Putative molecular mechanism of the estrogen positive feedback action on Kiss1 expression in the anteroventral-periventricular nucleus-periventricular nucleus continuum (AVPV-PeN). Preovulatory levels of circulating estrogen seem to act on AVPV-PeN kisspeptin neurons, in which estrogen-bound estrogen receptor α (ERα) may increase Kiss1 expression via histone acetylation of Kiss1 promotor region and chromatin loop formation between the Kiss1 promoter and the 3′-intergenic regions of Kiss1 locus. In the absence of estrogen, AVPV-PeN Kiss1 expression may be down-regulated by histone deacetylation of the Kiss1 promotor region and unwinding chromatin loops between the Kiss1 promoter and the 3′-intergenic regions of the Kiss1 locus.
Ijms 22 09229 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Uenoyama, Y.; Inoue, N.; Nakamura, S.; Tsukamura, H. Kisspeptin Neurons and Estrogen–Estrogen Receptor α Signaling: Unraveling the Mystery of Steroid Feedback System Regulating Mammalian Reproduction. Int. J. Mol. Sci. 2021, 22, 9229. https://doi.org/10.3390/ijms22179229

AMA Style

Uenoyama Y, Inoue N, Nakamura S, Tsukamura H. Kisspeptin Neurons and Estrogen–Estrogen Receptor α Signaling: Unraveling the Mystery of Steroid Feedback System Regulating Mammalian Reproduction. International Journal of Molecular Sciences. 2021; 22(17):9229. https://doi.org/10.3390/ijms22179229

Chicago/Turabian Style

Uenoyama, Yoshihisa, Naoko Inoue, Sho Nakamura, and Hiroko Tsukamura. 2021. "Kisspeptin Neurons and Estrogen–Estrogen Receptor α Signaling: Unraveling the Mystery of Steroid Feedback System Regulating Mammalian Reproduction" International Journal of Molecular Sciences 22, no. 17: 9229. https://doi.org/10.3390/ijms22179229

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop