scholarly journals Gonadotrophin-releasing hormone receptors in GtoPdb v.2021.3

2021 ◽  
Vol 2021 (3) ◽  
Author(s):  
Laura H. Heitman ◽  
Adriaan P. IJzerman ◽  
Craig A. McArdle ◽  
Adam J Pawson

GnRH1 and GnRH2 receptors (provisonal nomenclature [39], also called Type I and Type II GnRH receptor, respectively [85]) have been cloned from numerous species, most of which express two or three types of GnRH receptor [85, 84, 114]. GnRH I (p-Glu-His-Trp-Ser-Tyr-Gly-Leu-Arg-Pro-Gly-NH2) is a hypothalamic decapeptide also known as luteinizing hormone-releasing hormone, gonadoliberin, luliberin, gonadorelin or simply as GnRH. It is a member of a family of similar peptides found in many species [85, 84, 114] including GnRH II (pGlu-His-Trp-Ser-His-Gly-Trp-Tyr-Pro-Gly-NH2 (which is also known as chicken GnRH-II). Receptors for three forms of GnRH exist in some species but only GnRH I and GnRH II and their cognate receptors have been found in mammals [85, 84, 114]. GnRH1 receptors are expressed by pituitary gonadotrophs, where they mediate the effects of GnRH on gonadotropin hormone synthesis and secretion that underpin central control of mammalian reproduction. GnRH analogues are used in assisted reproduction and to treat steroid hormone-dependent conditions [58]. Notably, agonists cause desensitization of GnRH-stimulated gonadotropin secretion and the consequent reduction in circulating sex steroids is exploited to treat hormone-dependent cancers of the breast, ovary and prostate [58]. GnRH1 receptors are selectively activated by GnRH I and all lack the COOH-terminal tails found in other GPCRs. GnRH2 receptors do have COOH-terminal tails and (where tested) are selective for GnRH II over GnRH I. GnRH2 receptors are expressed by some primates but not by humans [88]. Phylogenetic classifications divide GnRH receptors into three [85] or five groups [129] and highlight examples of gene loss through evolution, with humans retaining only one ancient gene. The structure of the GnRH1 receptor in complex with elagolix has been elucidated [132].

2019 ◽  
Vol 2019 (4) ◽  
Author(s):  
Laura H. Heitman ◽  
Adriaan P. IJzerman ◽  
Craig A. McArdle ◽  
Adam J. Pawson

GnRH1 and GnRH2 receptors (provisonal nomenclature [35], also called Type I and Type II GnRH receptor, respectively [78]) have been cloned from numerous species, most of which express two or three types of GnRH receptor [78, 77, 107]. GnRH I (p-Glu-His-Trp-Ser-Tyr-Gly-Leu-Arg-Pro-Gly-NH2) is a hypothalamic decapeptide also known as luteinizing hormone-releasing hormone, gonadoliberin, luliberin, gonadorelin or simply as GnRH. It is a member of a family of similar peptides found in many species [78, 77, 107] including GnRH II (pGlu-His-Trp-Ser-His-Gly-Trp-Tyr-Pro-Gly-NH2 (which is also known as chicken GnRH-II). Receptors for three forms of GnRH exist in some species but only GnRH I and GnRH II and their cognate receptors have been found in mammals [78, 77, 107]. GnRH1 receptors are expressed by pituitary gonadotrophs, where they mediate the effects of GnRH on gonadotropin hormone synthesis and secretion that underpin central control of mammalian reproduction. GnRH analogues are used in assisted reproduction and to treat steroid hormone-dependent conditions [53]. Notably, agonists cause desensitization of GnRH-stimulated gonadotropin secretion and the consequent reduction in circulating sex steroids is exploited to treat hormone-dependent cancers of the breast, ovary and prostate [53]. GnRH1 receptors are selectively activated by GnRH I and all lack the COOH-terminal tails found in other GPCRs. GnRH2 receptors do have COOH-terminal tails and (where tested) are selective for GnRH II over GnRH I. GnRH2 receptors are expressed by some primates but not by humans [81]. Phylogenetic classifications divide GnRH receptors into three [78] or five groups [122] and highlight examples of gene loss through evolution, with humans retaining only one ancient gene.


2009 ◽  
Vol 297 (3) ◽  
pp. C591-C600 ◽  
Author(s):  
Ann R. Finch ◽  
Christopher J. Caunt ◽  
Stephen P. Armstrong ◽  
Craig A. McArdle

Gonadotropin-releasing hormone (GnRH) acts via seven transmembrane receptors to stimulate gonadotropin secretion. Sustained stimulation desensitizes GnRH receptor (GnRHR)-mediated gonadotropin secretion, and this underlies agonist use in hormone-dependent cancers. Since type I mammalian GnRHR do not desensitize, agonist-induced internalization and downregulation may underlie desensitization of GnRH-stimulated gonadotropin secretion; however, research focus has recently shifted to anterograde trafficking, with the finding that human (h)GnRHR are mostly intracellular. Moreover, there is little direct evidence for agonist-induced trafficking of hGnRHR, and whether or not type I mammalian GnRHR show agonist-induced internalization is controversial. Here we use automated imaging to monitor expression and internalization of hemagglutinin (HA)-tagged hGnRHRs, mouse (m) GnRHR, Xenopus (X) GnRHRs, and chimeric receptors (hGnRHR with added XGnRHR COOH tails, h.XGnRHR) expressed by adenoviral transduction in HeLa cells. We find that agonists stimulate downregulation and/or internalization of mGnRHR and XGnRHR, that GnRH stimulates trafficking of hGnRHR and can stimulate internalization or downregulation of hGnRHR when steps are taken to increase cell surface expression (addition of the XGnRHR COOH tail or pretreatment with pharmacological chaperone). Agonist effects on internalization (of h.XGnRHR) and downregulation (of hGnRHR and h.XGnRHR) were not mimicked by a peptide antagonist and were prevented by a mutation that prevents GnRHR signaling, demonstrating dependence on receptor signaling as well as agonist occupancy. Thus agonist-induced internalization and downregulation of type I mammalian GnRHR occurs in HeLa cells, and we suggest that the high throughput imaging systems described here will facilitate study of the molecular mechanisms involved.


Reproduction ◽  
2003 ◽  
pp. 271-278 ◽  
Author(s):  
AJ Pawson ◽  
K Morgan ◽  
SR Maudsley ◽  
RP Millar

Humans may be particularly unusual with respect to the gonadotrophin-releasing hormone (GnRH) control of their reproductive axis in that they possess two distinct GnRH precursor genes, on chromosomes 8p11-p21 and 20p13, but only one conventional GnRH receptor subtype (type I GnRH receptor) encoded within the genome, on chromosome 4. A disrupted human type II GnRH receptor gene homologue is present on chromosome 1q12. The genes encoding GnRH ligand precursors and GnRH receptors have now been characterized in a broad range of vertebrate species, including fish, amphibians and mammals. Ligand precursors and receptors can be categorized into three phylogenetic families. Members of each family exist in primitive vertebrates, whereas mammals exhibit selective loss of ligand precursor and receptor genes. One interpretation of these findings is that each ligand-cognate receptor family may have evolved to fulfil a separate function in reproductive physiology and that species-specific gene inactivation, modification or loss may have occurred during evolution when particular roles have become obsolete or subject to regulation by a different biochemical pathway. Evidence in support of this concept is available following the characterization of the chromosomal loci encoding the human type II GnRH receptor homologue, a rat type II GnRH receptor gene remnant (on rat chromosome 18) and a mouse type II GnRH ligand precursor gene remnant (on mouse chromosome 2). Whether type I GnRH and type II GnRH peptides elicit different signalling responses in humans by activation of the type I GnRH receptor in a cell type-specific fashion remains to be shown. Recent structure-function studies of GnRH ligands and GnRH receptors and their expression patterns in different tissues add further intrigue to this hypothesis by indicating novel roles for GnRH such as neuromodulation of reproductive function and direct regulation of peripheral reproductive tissues. Surprises concerning the complexities of GnRH ligand and receptor function in reproductive endocrinology should continue to emerge in the future.


1986 ◽  
Vol 108 (3) ◽  
pp. 441-449 ◽  
Author(s):  
C. A. Wilson ◽  
H. J. Herdon ◽  
L. C. Bailey ◽  
R. N. Clayton

ABSTRACT In the long-term castrated rat the negative feedback effect of testosterone is markedly reduced and the raised levels of plasma LH seen in the castrated animals are not suppressed by physiological concentrations of plasma testosterone. In this study we have measured pituitary gonadotrophin-releasing hormone (GnRH) receptor content as well as plasma and pituitary LH on days 1, 10 and 40 after castration and noted the effect of testosterone replacement on these parameters. We found that the negative feedback effect of physiological concentrations of testosterone on plasma and pituitary LH, pituitary GnRH receptor content and response to exogenous GnRH was attenuated 10 and 40 days after castration. It is suggested that the lack of effect of testosterone in the long-term castrated rat is due to its inability to reduce the pituitary GnRH receptor content. On increasing testosterone to supraphysiological levels, the negative feedback effect was reinstated. We also found that in rats 40 days after castration, physiological and subphysiological concentrations of testosterone significantly increased pituitary GnRH receptor content and this may explain the previous findings that low concentrations of testosterone can enhance the effect of GnRH and increase plasma LH levels. J. Endocr. (1986) 108, 441–449


2020 ◽  
Vol 27 (36) ◽  
pp. 6136-6158 ◽  
Author(s):  
Haralambos Tzoupis ◽  
Agathi Nteli ◽  
Maria-Eleni Androutsou ◽  
Theodore Tselios

Background: Gonadotropin-Releasing Hormone (GnRH) is a key element in sexual maturation and regulation of the reproductive cycle in the human organism. GnRH interacts with the pituitary cells through the activation of the Gonadotropin Releasing Hormone Receptors (GnRHR). Any impairments/dysfunctions of the GnRH-GnRHR complex lead to the development of various cancer types and disorders. Furthermore, the identification of GnRHR as a potential drug target has led to the development of agonist and antagonist molecules implemented in various treatment protocols. The development of these drugs was based on the information derived from the functional studies of GnRH and GnRHR. Objective: This review aims at shedding light on the versatile function of GnRH and GnRH receptor and offers an apprehensive summary regarding the development of different agonists, antagonists and non-peptide GnRH analogues. Conclusion: The information derived from these studies can enhance our understanding of the GnRH-GnRHR versatile nature and offer valuable insight into the design of new more potent molecules.


2003 ◽  
Vol 176 (2) ◽  
pp. 275-284 ◽  
Author(s):  
J Franklin ◽  
J Hislop ◽  
A Flynn ◽  
CA McArdle

Gonadotrophin-releasing hormone receptors (GnRH-Rs) are found in cancers of reproductive tissues, including those of the prostate, and gonadotrophin-releasing hormone (GnRH) can inhibit growth of cell lines derived from such cancers. Although pituitary and extra-pituitary GnRH-R transcripts appear identical, their functional characteristics may differ. Most extra-pituitary GnRH-Rs have low affinity for GnRH analogues and may not activate phospholipase C or discriminate between agonists and antagonists in the same way as do pituitary GnRH-Rs. Here we have assessed whether GnRH-Rs expressed exogenously in prostate cancer cells differ functionally from those of gonadotrophs. We found no evidence for endogenous GnRH-Rs in PC3 cells, but after infection with adenovirus expressing the GnRH-R (Ad GnRH-R) at 10 plaque forming units (p.f.u.)/cell or greater, at least 80% of the cells expressed GnRH-Rs. These sites had high affinity (K(d )for [(125)I]Buserelin 1.1+/-0.4 nM) and specificity (rank order of potency: Buserelin> GnRH>>chicken (c) GnRH-II), and mediated stimulation of [(3)H]inositol phosphate (IP) accumulation. Increasing viral titre from 3 to 300 p.f.u./cell increased receptor number (2000 to 275 000 sites/cell respectively) and [(3)H]IP responses. GnRH also caused a biphasic increase in the cytoplasmic Ca(2+) concentration in Ad GnRH-R-infected cells but not in control cells. Mobilization of Ca(2+) from intracellular stores contributed to the spike phase of this response whereas the plateau phase was dependent upon Ca(2+) entry across the plasma membrane. This effect on Ca(2+) and stimulation of [(3)H]IP accumulation were both blocked by the GnRH-R antagonist, Cetrorelix. In addition, GnRH reduced cell number (as measured in MTT activity assays) and DNA synthesis (as measured by [(3)H]thymidine incorporation) in Ad GnRH-R-infected cells (but not in control cells). This effect was mimicked by agonist analogues and inhibited by two antagonists. Thus, when exogenous GnRH-Rs are expressed at a density comparable to that in gonadotrophs, they are functionally indistinguishable from the endogenous GnRH-Rs in gonadotrophs. Moreover, expression of high affinity GnRH-Rs can facilitate a direct anti-proliferative effect of GnRH agonists on prostate cancer cells.


1993 ◽  
Vol 136 (1) ◽  
pp. 51-NP ◽  
Author(s):  
L. Anderson ◽  
G. Milligan ◽  
K. A. Eidne

ABSTRACT The present study has characterized the gonadotrophin-releasing hormone (GnRH) receptor in immortalized αT3-1 pituitary gonadotroph cells. GnRH and GnRH analogues produced both a dose- and time-dependent increase in total inositol phosphate (IP) accumulation. The rank order of potency of these analogues was the same as that obtained in parallel receptor-binding studies in αT3-1 cells. These responses were abolished following pretreatment with a GnRH antagonist. The use of a specific inositol 1,4,5-trisphosphate (Ins(1,4,5)P3) assay demonstrated a rapid but short-lived rise in Ins(1,4,5)P3 production. Intracellular calcium ([Ca2+]i) was subsequently measured in αT3-1 cells using dual wavelength fluorescence microscopy combined with dynamic video imaging. GnRH produced a biphasic rise in [Ca2+]i. The initial calcium transient was complete within seconds while the smaller secondary plateau phase lasted several minutes. G-protein involvement in the IP response to GnRH in αT3-1 cells was investigated using sodium fluoride (NaF) and pertussis toxin (PTx) which activate and inactivate G-proteins respectively. Like GnRH, NaF produced a dose- and time-dependent increase in IP accumulation. Activation of phospholipase C in these cells by either GnRH or NaF was PTx-insensitive, suggesting that the G-protein involved was neither Gi nor Go but more probably Gq. Immunoblot analysis of αT3-1 cell membranes using antisera raised against the predicted C-terminal decapeptide of the α subunit of Gq demonstrated the presence of Gq in αT3-1 cells. Collectively these results show that the GnRH receptors expressed in αT3-1 cells are coupled to the phosphatidylinositol second messenger pathway via a specific G-protein. αT3-1 therefore represents a convenient model in which to study GnRH-related second messenger pathways. Journal of Endocrinology (1993) 136, 51–58


1992 ◽  
Vol 8 (2) ◽  
pp. 109-118 ◽  
Author(s):  
J. Brooks ◽  
W. J. Crow ◽  
J. R. McNeilly ◽  
A. S. McNeilly

ABSTRACT The modulation of FSH secretion at the beginning and middle of the follicular phase of the cycle represents the key event in the growth and selection of the preovulatory follicle. However, the mechanisms that operate within the pituitary gland to control the increased release of FSH and its subsequent inhibition in vivo remain unclear. Treatment of ewes with bovine follicular fluid (bFF) during the luteal phase has been previously shown to suppress the plasma concentrations of FSH and, following cessation of treatment on day 11, a rebound release of FSH occurs on days 12 and 13. When luteal regression is induced on day 12, this hypersecretion of FSH results in an increase in follicle growth and ovulation rate. To investigate the mechanisms involved in the control of FSH secretion, ewes were treated with twice daily s.c. injections of 5 ml bFF on days 3–11 of the oestrous cycle and luteal regression was induced on day 12 with prostaglandin (PG). The treated ewes and their controls were then killed on day 11 (luteal), or 16 or 32h after PG and their pituitaries removed and halved. One half was analysed for gonadotrophin and gonadotrophin-releasing hormone (GnRH) receptor content. Total pituitary RNA was extracted from the other half and subjected to Northern analysis using probes for FSH-β, LH-β and common α subunit. Frequent blood samples were taken and assayed for gonadotrophins. FSH secretion was significantly (P<0.01) reduced during bFF treatment throughout the luteal phase and then significantly (P<0.01) increased after cessation of treatment, with maximum secretion being reached 18– 22h after PG, and then declining towards control values by 32h after PG. A similar pattern of LH secretion was seen after bFF treatment. Pituitary FSH content was significantly (P<0.05) reduced by bFF treatment at all stages of the cycle. No difference in the pituitary LH content was seen. The increase in GnRH receptor content after PG in the controls was delayed in the treated animals. Analysis of pituitary mRNA levels revealed that bFF treatment significantly (P<0.01) reduced FSH-β mRNA levels in the luteal phase. Increased levels of FSH-β, LH-β and α subunit mRNA were seen 16h after PG in the bFF-treated animals, at the time when FSH and LH secretion from the pituitary was near maximum. These results suggest that the rebound release of FSH after treatment with bFF (as a source of inhibin) is related to a rapid increase in FSH-β mRNA, supporting the concept that the rate of FSH release is directly related to the rate of synthesis.


2011 ◽  
Vol 23 (6) ◽  
pp. 780 ◽  
Author(s):  
Magdalena Ciechanowska ◽  
Magdalena Łapot ◽  
Tadeusz Malewski ◽  
Krystyna Mateusiak ◽  
Tomasz Misztal ◽  
...  

There is no information in the literature regarding the effect of corticotropin-releasing hormone (CRH) on genes encoding gonadotrophin-releasing hormone (GnRH) and the GnRH receptor (GnRHR) in the hypothalamus or on GnRHR gene expression in the pituitary gland in vivo. Thus, the aim of the present study was to investigate, in follicular phase ewes, the effects of prolonged, intermittent infusion of small doses of CRH or its antagonist (α-helical CRH 9-41; CRH-A) into the third cerebral ventricle on GnRH mRNA and GnRHR mRNA levels in the hypothalamo–pituitary unit and on LH secretion. Stimulation or inhibition of CRH receptors significantly decreased or increased GnRH gene expression in the hypothalamus, respectively, and led to different responses in GnRHR gene expression in discrete hypothalamic areas. For example, CRH increased GnRHR gene expression in the preoptic area, but decreased it in the hypothalamus/stalk median eminence and in the anterior pituitary gland. In addition, CRH decreased LH secretion. Blockade of CRH receptors had the opposite effect on GnRHR gene expression. The results suggest that activation of CRH receptors in the hypothalamus of follicular phase ewes can modulate the biosynthesis and release of GnRH through complex changes in the expression of GnRH and GnRHR genes in the hypothalamo–anterior pituitary unit.


Sign in / Sign up

Export Citation Format

Share Document