ArticlePDF AvailableLiterature Review

Relaxin Family Peptides and Their Receptors

Authors:

Abstract

There are seven relaxin family peptides that are all structurally related to insulin. Relaxin has many roles in female and male reproduction, as a neuropeptide in the central nervous system, as a vasodilator and cardiac stimulant in the cardiovascular system, and as an antifibrotic agent. Insulin-like peptide-3 (INSL3) has clearly defined specialist roles in male and female reproduction, relaxin-3 is primarily a neuropeptide involved in stress and metabolic control, and INSL5 is widely distributed particularly in the gastrointestinal tract. Although they are structurally related to insulin, the relaxin family peptides produce their physiological effects by activating a group of four G protein-coupled receptors (GPCRs), relaxin family peptide receptors 1-4 (RXFP1-4). Relaxin and INSL3 are the cognate ligands for RXFP1 and RXFP2, respectively, that are leucine-rich repeat containing GPCRs. RXFP1 activates a wide spectrum of signaling pathways to generate second messengers that include cAMP and nitric oxide, whereas RXFP2 activates a subset of these pathways. Relaxin-3 and INSL5 are the cognate ligands for RXFP3 and RXFP4 that are closely related to small peptide receptors that when activated inhibit cAMP production and activate MAP kinases. Although there are still many unanswered questions regarding the mode of action of relaxin family peptides, it is clear that they have important physiological roles that could be exploited for therapeutic benefit.
RELAXIN FAMILY PEPTIDES
AND THEIR RECEPTORS
R. A. D. Bathgate, M. L. Halls, E. T. van der Westhuizen, G. E. Callander, M. Kocan,
and R. J. Summers
Drug Discovery Biology, Monash Institute of Pharmaceutical Sciences & Department of Pharmacology, Monash
University, Victoria, Australia; Florey Neurosciences Institutes and Department of Biochemistry and Molecular
Biology, The University of Melbourne, Victoria, Australia; and Université de Montréal, Institut de Recherche en
Immunologie et Cancérologie, Pharmacologie Moléculaire, Montréal, Canada
LBathgate RAD, Halls ML, van der Westhuizen ET, Callander GE, Kocan M,
Summers RJ. Relaxin Family Peptides and Their Receptors. Physiol Rev 93: 405–
480, 2013; doi:10.1152/physrev.00001.2012.—There are seven relaxin family
peptides that are all structurally related to insulin. Relaxin has many roles in female
and male reproduction, as a neuropeptide in the central nervous system, as a
vasodilator and cardiac stimulant in the cardiovascular system, and as an antifibrotic agent.
Insulin-like peptide-3 (INSL3) has clearly defined specialist roles in male and female reproduc-
tion, relaxin-3 is primarily a neuropeptide involved in stress and metabolic control, and INSL5
is widely distributed particularly in the gastrointestinal tract. Although they are structurally
related to insulin, the relaxin family peptides produce their physiological effects by activating a
group of four G protein-coupled receptors (GPCRs), relaxin family peptide receptors 1–4
(RXFP1–4). Relaxin and INSL3 are the cognate ligands for RXFP1 and RXFP2, respectively,
that are leucine-rich repeat containing GPCRs. RXFP1 activates a wide spectrum of signaling
pathways to generate second messengers that include cAMP and nitric oxide, whereas RXFP2
activates a subset of these pathways. Relaxin-3 and INSL5 are the cognate ligands for RXFP3
and RXFP4 that are closely related to small peptide receptors that when activated inhibit cAMP
production and activate MAP kinases. Although there are still many unanswered questions
regarding the mode of action of relaxin family peptides, it is clear that they have important
physiological roles that could be exploited for therapeutic benefit.
I. INTRODUCTION: THE DISCOVERY OF... 405
II. MOLECULAR BIOLOGY, SYNTHESIS,... 413
III. MOLECULAR BIOLOGY, STRUCTURAL... 417
IV. EXPRESSION AND SECRETION OF... 425
V. LOCALIZATION OF RELAXIN FAMILY... 429
VI. SIGNAL TRANSDUCTION PATHWAYS... 431
VII. PHYSIOLOGICAL ROLES OF... 443
VIII. POTENTIAL THERAPEUTIC... 456
IX. MAJOR UNANSWERED QUESTIONS... 459
I. INTRODUCTION: THE DISCOVERY OF
RELAXIN FAMILY PEPTIDES AND THEIR
RECEPTORS
A. The Discovery of the Relaxin Family
Peptides
Relaxin was discovered and named by Dr. Frederick Hisaw
following observations of the reproductive endocrinology
of the gopher and later the guinea pig. He noticed that there
was softening and expansion of the pubic ligament just
prior to delivery in the pregnant female that facilitated par-
turition. In 1926 he showed that the injection of serum from
pregnant guinea pigs or rabbits caused relaxation of the
pubic ligament of virgin guinea pigs when given shortly
after estrus (220). The following year the same relaxing
factor was shown to be present in pig corpus luteum and
rabbit placenta (221). The hormone was formally named
relaxin after it was extracted from pig corpus luteum in
1930 (165). For the next 15 years or so, relatively little
work was done with relaxin, but post World War II there
was a reawakening of interest in its physiological role that
established properties that would be useful in understand-
ing its roles in pregnancy and parturition. These included,
in estrogen-primed animals, expansion of the public liga-
ment of mice (193, 195), relaxation of the uterine myome-
trium in guinea pigs (295), and cervical softening in cattle.
The subsequent purification of relaxin from animal sources
(initially pigs but then many other species) led to the deter-
mination of the peptide structure, biological actions and
generation of reliable bioassays. The information obtained
from the determination of the peptide structure enabled the
use of recombinant DNA techniques to clone the rat (244)
and pig (192) relaxin cDNAs that demonstrated a close
structural relationship with insulin (FIGURE 1A). Screening
Physiol Rev 93: 405–480, 2013
doi:10.1152/physrev.00001.2012
4050031-9333/13 Copyright © 2013 the American Physiological Society
of a human genomic library resulting in the cloning of hu-
man gene-1 relaxin (RLN1) (245) followed by the isolation
of a second gene, human gene-2 relaxin (RLN2), from a
human luteal cDNA library (246). It has subsequently been
shown that humans and higher primates have RLN1 and
RLN2 genes, whereas other mammals have only RLN1.
Importantly, the peptide encoded by the human RLN2 gene
and the RLN1 gene from other mammals encodes the re-
laxin peptide circulating during pregnancy. The function of
the RLN1 gene in humans and higher primates is unknown.
The ability to clone human relaxin genes and species ho-
mologs led to the production of relaxin by recombinant
techniques and the development of the relaxin knockout
A B
B-chains:
A-chains:
INSL3
INSL6
INS
IGF2
IGF1
INSL4 INSL5
RLN2
RLN1
RLN3
R3
R2
I3
I3
I5
I5
R3
R2
I5
I3
I3
I5
R3 R2
R3
R2
CysB10
CysB22
CysA24
CysA10
CysA15
CysA15
CysA11
CysA11
CysA11 CysA15
A-chains B-chains
C
FIGURE 1. Relationship between relaxin family peptides, insulin, and insulin-like peptides. A: CLUSTALW
(v1.83) alignment of human relaxin/insulin family peptides. Relaxin and insulin family peptide sequences
are from the UniProt/SwissProt database: human relaxin-1 (P04808), human relaxin-2 (P04090),
human relaxin-3 (Q8WXF3), human INSL3 (P51460), human INSL4 (Q14641), human INSL5 (Q9Y5Q6),
human INSL6 (Q9Y581), human IGFI (P01343), human IGFII (P01344), and human insulin (P01308).
Conserved cysteine residues are black with gray background. Solid black lines indicate disulfide bonds.
Identical amino acid residues are indicated by *. The insulin-like growth factors are not cleaved from the
pro-hormone state, thus indicating that the sequence of these peptides is continuous. B: an unrooted
phylogenetic tree derived using the neighbor joining method in CLUSTALW. Human peptide sequences
were edited to delete the C and signal peptide sequences. The resulting domains corresponding to the B
and A chains were further edited to minimize gaps, and then concatenated (562). The phylogenetic tree
was formatted using the program Dendroscope (249). C: A-chains and B-chains of the cognate relaxin
family peptides for RXFP1–4 superimposed on one another. This demonstrates the extraordinary degree
of structural similarity between the peptides that differ mainly in the mid and COOH-terminal regions of
the B-chain. R2 (magenta), human relaxin-2; R3 (blue), human relaxin-3; I3 (red), INSL3; I5 (green),
INSL5.
BATHGATE ET AL.
406 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
mouse. Furthermore, cloning strategies or database search-
ing using the conserved relaxin peptide structure led to the
discovery of a variety of related peptides. These include
insulin-like peptide 3 (INSL3) (2), INSL4 (94) by differen-
tial cloning, and INSL5 (100) and INSL6 (236) by ex-
pressed sequence tags (EST) database screening. The most
recent addition, relaxin-3, was discovered by screening the
human genomic database (46). In humans, the seven relaxin
family peptides that have been identified are encoded by
genes that possess a similar structure (FIGURE 1, AAND C)
and synthesize similar prepropeptides (241). The mature
peptides are all predicted or proven to be processed from
the prepropeptides by convertases to produce two chain
peptides that all have an A-chain linked to a B-chain by two
disulfide bonds with an additional intrachain disulfide in
the A chain (see section II).
Although many of the relaxin family peptides were discov-
ered as a result of their roles in reproduction, it is now clear
that they perform a wide variety of physiological functions
in humans. Some such as relaxin have widely differing roles
between species, whereas others such as relaxin-3 and
INSL3 have well conserved roles in all mammalian species.
B. The Relaxin Peptides
As mentioned above, a relaxin gene is present in all mam-
mals and is responsible for the production of the relaxin
peptide that is circulating in the blood during pregnancy.
However, the levels of relaxin in the blood during preg-
nancy are very different between species and correlate with
the varied roles for relaxin in pregnancy and parturition in
different species (see sect. IVA). In humans, blood levels of
relaxin are highest in the first trimester, and consequently,
relaxin is not associated with parturition in humans. Nev-
ertheless, it is likely that in humans, relaxin has a role in the
early stages of pregnancy and also in the cardiovascular
changes that accompany pregnancy. Furthermore, it is be-
coming increasingly clear that relaxin is produced in many
tissues in mammals as a paracrine or autocrine factor to
exert many different physiological roles (see sect. IVA).
Relaxin-3 is the most recently identified relaxin family pep-
tide and was discovered following a search of the human
genomic database (46). It is classified as a relaxin peptide
based on the RxxxRxxI/V relaxin-binding motif in the B-
chain, despite an otherwise relatively low sequence homol-
ogy to other relaxin peptides (FIGURE 1A). Unlike the other
relaxins, which show considerable species heterogeneity
apart from the relaxin binding motif and key structural
residues, the relaxin-3 sequences are well-conserved across
species (562, 574). It is now widely accepted that relaxin-3
is the ancestral peptide of the relaxin family peptides (562)
and that in mammals is primarily a neuropeptide expressed
almost exclusively in brain (46). Functions likely to be as-
sociated with relaxin-3 include stress, memory, and appe-
tite regulation (see sect. VIIC) (25, 329, 358, 516).
C. INSL3, 4, 5, and 6
INSL3 (formerly Leydig insulin-like peptide) was discov-
ered in the Leydig cells of the testis (2) and is highly ex-
pressed in these cells in all species that possess the INSL3
gene (47). Although there is evidence for INSL3 expression
in other tissues (see sect. IVB), this occurs at much lower
levels than in the Leydig cells. As such, INSL3 has a critical
role in testis descent, and INSL3 knockout mice are cryp-
torchid and as a consequence infertile (379, 596). INSL3
plays an important role in the development of the guber-
naculum, which is important in the first stage of testis de-
scent, and it also appears to have a role in the maintenance
of fertility in females (490) (see sect. VIIB).
Compared with the other relaxin family peptides, very little
information is available regarding INSL46. INSL4 is highly
expressed in the placenta (94) and may play a role in bone
development (303), but its target receptor has yet to be iden-
tified. INSL5 on the other hand has quite a wide distribution
pattern (see sect. IVD) with particularly high expression in the
gastrointestinal tract (100). Although INSL5 knockout mice
have been developed, the associated phenotype is currently not
well characterised; nevertheless, there may be functional ef-
fects on fat and glucose metabolism [patent WO2005124361
(A3)]. This is of interest given that the related peptides insulin
and the insulin-like growth factors, IGF-I and IGF-II, have
well-defined roles in glucose metabolism. INSL5 activates
RXFP4 with high potency but unlike relaxin-3 does not acti-
vate RXFP1 or RXFP3 and in fact is a weak antagonist at
RXFP3 (315). It is now widely accepted that INSL5 is the
cognate ligand for RXFP4 (212, 315). INSL6 was discovered
by screening EST databases (236, 275, 317) and like INSL3 is
found predominantly in the testis. Specific localization of pep-
tide expression in spermatocytes and spermatids suggests a
role for INSL6 in male fertility, and male INSL6 knockout
mice show a marked reduction in sperm numbers and motility
(85). The receptor for INSL6 is currently unknown.
D. Receptors for the Relaxins and
Insulin-like Peptides
The search for the cognate receptors for the relaxin family
peptides proved demanding and was complicated by the close
similarity between relaxin family peptides and insulin which
together with studies that showed an increase in tyrosine phos-
phorylation of a 220-kDa protein in response to relaxin (407)
suggested that they were tyrosine kinases. However, the key
insights came from an unlikely direction when it was noticed
that INSL3 knockout mice (379, 596) shared the same pheno-
type of abnormal testis descent as mice with disruptions in the
G protein-coupled receptor encoded by the GREAT gene
(later shown to be the mouse ortholog of human LGR8) (406).
This suggested that the targets for relaxin family peptides
might be GPCRs and subsequently led to the deorphanization
of LGR7 and LGR8 (239). Construction of constitutively ac-
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
407Physiol Rev VOL 93 JANUARY 2013 www.prv.org
tive mutants of LGR7 and LGR8 showed that these receptors
likely signaled by activating the adenylyl cyclase pathway and
further stimulation of these receptors by relaxin resulted in
increases in cAMP (239). A subsequent study demonstrated
that, as anticipated from the studies above, LGR8 is the recep-
tor for INSL3 (65, 301). Two other orphan receptors
GPCR135 and GPCR142 were shown to be the targets for
relaxin-3 and INSL5, respectively (313, 314). These receptors
have now been classified as relaxin family peptide (RXFP)
receptors: RXFP1 (LGR7*), RXFP2 (LGR8*), RXFP3
(GPCR135*), and RXFP4 (GPCR142*) (*these are the previ-
ously used names for the RXFP receptors prior to reclassifica-
tion by NC-IUPHAR) (FIGURE 2D)(42). Cognate receptors for
INSL4 and INSL6 have yet to be identified, and these peptides
do not interact with any of the RXFP receptors.
RXFP1 is the cognate receptor for human relaxin-2 in humans
that in addition to the classical 7-transmembrane spanning
regions has a large extracellular domain containing 10 leucine-
rich repeats (LRR) and a unique NH
2
-terminal low-density
lipoprotein receptor type A (LDLa) module (239) (FIGURE 3).
As befitting the roles of relaxin in reproduction, RXFP1
mRNA and protein are found in a wide range of reproductive
tissues including ovary, uterus, placenta, mammary gland,
prostate, and testis (FIGURE 4). The receptor is also found in
the heart, kidney, lung, liver, and blood cells as well as in a
number of areas of brain such as cortex, organum vasculosum
of the lamina terminalis (OVLT), and subfornical organ (SFO)
(see TABLE 1 and sect. VA). Thus relaxin not only has auto-
crine and paracrine roles but also acts as a neuropeptide. In-
teraction of relaxin with RXFP1 involves at least three stages:
high-affinity binding between the -chain and the LRR region;
lower affinity binding to the transmembrane exoloops, and
finally an interaction involving the LDLa module that is essen-
tial for signaling (199, 501) (FIGURE 3). Although RXFP1 cou-
ples to numerous signal transduction pathways (see sect. VIA),
many early studies indicated that treatment with relaxin
caused increases in cAMP levels in tissues (69, 95, 96, 446).
This link with cAMP signaling was strengthened in the studies
that deorphanized LGR7 (RXFP1) where it was shown that a
constitutively active receptor mutant generated cAMP (239)
(FIGURE 3). This has been confirmed in numerous studies that
show that RXFP1 couples to at least three G proteins to induce
a complex pattern of cAMP production (198, 202, 204). In
addition, RXFP1 is known to activate Erk1/2, tyrosine ki-
nase(s), gene transcription, and nitric oxide (NO) signaling
and can also interact with the glucocorticoid receptor (GC).
The full implications of the pleiotropic effects of relaxin are
still to be elucidated.
RXFP2 shares very close structural similarity with RXFP1
(239) and is the cognate receptor for INSL3 (301). It has a
much more restricted distribution, physiological function,
and signaling profile than RXFP1 (FIGURE 4;TABLE 1).
RXFP2 is expressed in the gubernaculum as anticipated
from its critical role in testicular descent but is also ex-
pressed in the testis and ovary relating to roles in gonadal
function (276) and in bone relating to a role in bone metab-
olism (161) (see TABLE 1 and sect. VIIB). RXFP2 signaling
also involves adenylyl cyclase activation, but this is re-
FIGURE 2. Evolutionary relationships between human RXFP receptors and other family A GPCRs. Receptors showing similarity to RXFP1 or
RXFP3 were identified by using complete transmembrane domain amino acid sequences (from the beginning of helix 1 through to the COOH
terminus) to conduct BLAST searches against the human REFSEQ protein database. Sequences showing the greatest similarity with each of
RXFP1 or RXFP3 were then aligned using CLUSTALW, and a phylogenetic tree derived by the neighbor-joining method, also in CLUSTALW.
A: the phylogenetic tree shows that RXFP1 and RXFP2 are closely related to each other, with 59% sequence identity and 80% similarity based
on conservative amino acid changes. They form a subgroup that also includes the three glycoprotein hormone receptors for thyroid stimulating
hormone (TSHR), luteinizing hormone/chorionic gonadotropin (LHCGR), and follicle stimulating hormone (FSHR). These receptors show
somewhat higher homology to RXFP1 and RXFP2 (27–29% identity, 49–51% similarity) than a second group comprising the leucine-rich
repeat-containing receptors LGR4, LGR5, and LGR6 (22–25% identity, 46–47% similarity). RXFP1 and RXFP2 are also related to family A
bioamine receptors such as
-adrenoceptors (ARs) and adenosine receptors, as well as additional peptide receptors. B: the RXFP3 and RXFP4
amino acid sequences that display 43% sequence identity and 60% similarity are most closely related to a subgroup comprising the angiotensin
II type 1 receptor (AGTR1), the apelin receptor (APLNR) and the chemokine receptor (CCR1), and a second subgroup containing neuropeptide
receptors including members of the somatostatin and opioid receptor families and the neuropeptide B/W receptor (NPBWR1). They are also related
to various chemokine receptors (CXCR7, CXCR4, CCR1, and CCR9) and to the formyl peptide receptors (FPR2 and FPR3), C: the RXFP1/RXFP2 and
RXFP3/RXFP4 receptor subgroups are not highly related to each other, and in fact, the only significant homology between RXFP1 and RXFP3 is
within the highly conserved TM7 motif NSxLNP(I/V)Y that is essential for G protein coupling as well as neighboring key residues in helix 8. RXFP1
and RXFP4 show slightly higher homology, although again the regions of similarity encompass only short motifs in TM3, TM6, and TM7. Instead,
both receptor subgroups are related to other family A GPCRs such as the
2
-AR and the adenosine A2B receptor. In essence, these bioamine
receptors occupy a bridge between the distinct RXFP1/RXFP2 and RXFP3/RXFP4 subgroups. For example, the
2
-AR has homologies with
each of the RXFP1 and RXFP3 receptors, with 23% identity/39% similarity to RXFP1 and 21% identity/38% similarity to RXFP3. D: RXFP1
and RXFP2 are closely related to each other and have a characteristic extracellular leucine-rich repeat region and a unique low-density lipoprotein
receptor class A module at the NH2 terminus. In contrast, RXFP3 and RXFP4 are both similar to small peptide (e.g., bradykinin, angiotensin
II) liganded receptors. The primary ligands for each RXFP are shown in FIGURE 1. Whereas the RXFP1/RXFP2 and RXFP3/RXFP4 receptors
belong to separate subtrees, the endogenous ligands do not show a corresponding split in lineage. Based on BLASTP and also PSI-BLAST
searches using the human relaxin-2 B chain sequence, human relaxin-3 is somewhat more similar than INSL3 (57% identity, 67% similarity
versus 52% identity, 62% similarity, respectively), while INSL5 is the least related. Searches using the A chain sequence indicate that human
relaxin-3 and INSL3 are equivalent in their homology to human relaxin-2, and again INSL5 is more distantly related. Thus human relaxin-3 is
closely related to human relaxin-2 despite the corresponding primary receptors residing in different subgroups. This is not surprising given the
distinct nature of the ligand binding domains in each receptor subgroup and indicates that the capacity to bind related peptides has arisen by
a process of convergent evolution.
BATHGATE ET AL.
408 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
SSTR1
SSTR5
SSTR2
RXFP4
RXFP3
RXFP1
RXFP2
TSHR
FSHR
LHCGR
LGR5
LGR6
LGR4
APLNR
CCR1
AGTR1
OPRL1
OPRM1 NPBWR1 ADRB2
ADORA2B
NPFF2 RXFP1 RXFP2
TSHR
LHCGR
FSHR
LGR5
LGR6
LGR4
GHSR
OXTR
SSTR1
SSTR4
AGTR1
AGTR2
ADORA2B
A
DORA2A
ADORA1
ADRB1
ADRB2
SSTR1
SSTR5
SSTR2
SSTR4
OPRM1
OPRL1
OPRK1
OPRD1
NPBWR1
CCR1
CXCR4
CCR9
FPR3FPR2
CXCR7
AGTR2
AGTR1
APLNR RXFP3 RXFP4
Extracellular
Intracellular
RXFP3/4 RXFP1/2
AB
C
D
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
409Physiol Rev VOL 93 JANUARY 2013 www.prv.org
stricted to a subset of the G proteins utilized by RXFP1 (see
sect. VIB). Although in the cell systems often used to study
RXFP2 activation of the receptor causes an overwhelming
increase in cAMP levels, in endogenously expressing sys-
tems both increases and decreases in cAMP are seen. For
instance, in gubernacular cells (301) or osteoblasts (161),
activation of RXFP2 causes increased cAMP, but in male
germ cells and oocytes, the opposite effect is observed (276).
This may reflect the particular expression pattern of signal-
ing proteins in the various cells (202). Interestingly, some
species relaxins have been shown to activate RXFP2 in vitro
(45, 239, 301, 453) but the physiological significance (if
any) of this interaction is not clear. There is no evidence that
suggests that relaxin can activate RXFP2 in vivo. Mouse
relaxin does not activate RXFP2 (45), indicating that stud-
ies in the RXFP1 knockout mouse will not resolve possible
physiological actions of relaxin at RXFP2.
RXFP3 and RXFP4 are both similar to type 1 small peptide
receptors and are distinctly different in structure from
RXFP1 and RXFP2 (FIGURE 2D). RXFP3 was originally
named SALPR or somatostatin and angiotensin-like pep-
Leucine-rich repeat region
Primary high affinity binding site
LDLa module
LDLa-less receptor binds but
does not signal
ECL region
Secondary lower affinity
binding site
IL3-Gαs coupling
Asp637 constitutive activity
Helix 8 Interaction
with AKAP79
C-tail Ser704 β-arrestin-2 binding
Final 10 residues + Arg752 - Gαi3 activation
RXFP1
FIGURE 3. Structural features of the relaxin family peptide receptor RXFP1. The RXFP1 receptor has a large
extracellular domain that contains an NH
2
-terminal low-density lipoprotein receptor type A (LDLa) module
connected to a leucine-rich repeat (LRR) region tethered to the transmembrane domains of the receptor by a
unique hinge region (237). The primary high-affinity binding site for relaxin is located in the LRR region, but the
peptide also interacts with a lower affinity binding site located on the extracellular loops (199, 501). An intact
LDLa module at the NH
2
terminus is essential for signaling (231, 280, 456) which involves the third
intracellular loop. Protein-protein interactions important for signaling occur with helix 8 and the COOH-terminal
tail of RXFP1 (196, 198, 201, 204).
BATHGATE ET AL.
410 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
tide receptor (352). Similarly, RXFP4 was originally
thought to be a bradykinin receptor (63) before both
RXFP3 and RXFP4 were paired with their cognate ligands
relaxin-3 (314) and INSL5 (315). Importantly, relaxin-3
will bind to and activate RXFP4 as well as RXFP1; in ad-
dition, INSL5 is a low-affinity antagonist of RXFP3. The
receptor expression profiles (FIGURE 4) match potential
roles as a neuropeptide receptor in the case of RXFP3 and a
gut hormone receptor in the case of RXFP4 (see sect. VII, C
and D). Relatively little is known of the signaling profiles
exhibited by RXFP3 (see sect. VIC) and RXFP4 (see sect.
VID). RXFP3 couples to G
i/o
and inhibits adenylyl cyclase
(314, 546), and receptor stimulation also activates ERK1/2
phosphorylation (546). Although earlier studies indicated
that only relaxin-3 and its -chain could activate RXFP3,
more recent work suggests that relaxin can also activate
specific pathways by binding to a subtly different site on
RXFP3 (545). Understanding this ligand-directed signaling
bias will be important for the development of RXFP3 as a
therapeutic target.
Relaxin-RXFP1 Relaxin-3-RXFP3
INSL3-RXFP2 INSL5-RXFP4
Brain
Heart
Vasculature
Breast
Lung
Skin
Kidney
Liver Gut
Bone
Uterus
Ovary
Testis
Prostate
FIGURE 4. Expression patterns of relaxin family peptides and their cognate RXFP receptors in the major
organ systems of the body. Although there appears to be in some cases overlap between the distribution of
ligand/receptor pairs and in some cases there has been shown to be interaction between particular peptides
and several RXFP receptors in vitro, this does not appear to occur in vivo. Thus human relaxin-2 interacts with
RXFP1 and RXFP2 and human relaxin-3 interacts with RXFP1, RXFP3, and RXFP4 in vitro, but there is no
evidence for such interactions occurring in a physiological setting.
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
411Physiol Rev VOL 93 JANUARY 2013 www.prv.org
Table 1. Tissue localization of RXFP1-4 receptors
RXFP1 RXFP2 RXFP3 RXFP4
Tissue/Species Rat Mouse Human Rat Mouse Human Rat Mouse Human Human
Ovary mRNA
1
Protein
23
mRNA
2,54
mRNA
17
Oviduct mRNA
1
mRNA
3
Protein
35
Uterus mRNA
1
mRNA
3,14,30
mRNA
2,4,8
mRNA
2
Protein
2,20,21
Protein
23
Protein
4,8,25
Uterine smooth
muscle mRNA
38,46
mRNA
3
Protein
25
Protein
20,21,38,46
Protein
2
Endometrium mRNA
34
Protein
2
mRNA
10,18,19,49
Protein
18,19
mRNA
10
Cervix, vagina Protein
2,20,22
mRNA
3,30
Protein
25
Protein
2
Placenta mRNA
30
mRNA
2,19
mRNA
39
mRNA
37
Nipple Protein
22
mRNA
3,30
Protein
25
Breast Protein
22
Protein
4,25
mRNA
39
Testis mRNA
1,6,51
mRNA
3,15,30,47
mRNA
2
mRNA
1,17,51
mRNA
11
mRNA
2
mRNA
47
mRNA
36,39
mRNA
37
Protein
51
Protein
51
Prostate mRNA
50
mRNA
15
mRNA
2
mRNA
48
mRNA
37
Gubernaculum mRNA
6,12
mRNA
11
Brain mRNA
1,5,6,9,51
mRNA
3,30
mRNA
2
mRNA
31
mRNA
11
mRNA
2
Protein
40
mRNA
36,39
mRNA
37
Protein
20,21
Protein
23
Brain regions mRNA
27
mRNA
28,29
mRNA
33,29,31
mRNA
29
mRNA
41,42,52
mRNA
4345,53
mRNA
36,39
Protein
27
Protein
28,29
Protein
33,29
Protein
29
Protein
41,42,52
Protein
44,53
Pituitary mRNA
3
mRNA
39
Kidney mRNA
1
mRNA
2
mRNA
32
mRNA
2
mRNA
37
Heart mRNA
1,5,9,51
mRNA
3,30
mRNA
2
Protein
21,24
Lung mRNA
7,30
Liver
Intestine mRNA
1
mRNA
30
Colon mRNA
1
mRNA
37
Pancreas mRNA
36,39
Adrenal mRNA
1
mRNA
2
mRNA
36,39
Thyroid mRNA
2,16
mRNA
37
Thymus mRNA
36
mRNA
37
Salivary glands mRNA
36,39
mRNA
37
Muscle mRNA
2
Blood cells mRNA
2
Continued
BATHGATE ET AL.
412 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
II. MOLECULAR BIOLOGY, SYNTHESIS,
STRUCTURE, AND IDENTIFICATION OF
FEATURES IMPORTANT FOR ACTIVITY
OF RELAXIN FAMILY PEPTIDES
A. Molecular Biology of Relaxin Family
Peptides
Although the relaxin peptide was initially described in the
1920s, the first relaxin gene sequence was not cloned until
the early 1980s. At this time, cloning of porcine (192) and
rat (244) relaxin was achieved by screening of corpus lu-
teum cDNA libraries. These studies confirmed that relaxin,
like insulin, is synthesized as a preprohormone with four
distinct regions: a signal peptide, a -chain, a C-chain, and
a COOH-terminal A-chain. Relaxin cDNA and gene se-
quences were subsequently cloned from a number of mam-
malian species (43). Additionally, multiple new mammalian
relaxin gene sequences are now available in the genome
databases. All of these sequences show a similar gene struc-
ture and possess two exons interrupted by an intron, which
is situated in the middle of the C-peptide sequence.
The first human relaxin gene was cloned by screening a
genomic library using the porcine relaxin cDNA sequence
and was designated the human RLN1 gene with the peptide
product termed H1 relaxin (human relaxin-1) (245).
1
How-
ever, a year later, the same researchers isolated an addi-
tional relaxin gene by screening a corpus luteum cDNA
library (246). This gene was designated RLN2 and the gene
product termed H2 relaxin (human relaxin-2) (see footnote 1).
The RLN2 product sequence was identical to that of the
relaxin peptide that was subsequently isolated from the hu-
man corpus luteum (126, 566), and two relaxin genes are
only present in higher primates (562). Thus the human re-
laxin-2 peptide is the equivalent of the relaxin peptide first
discovered by Hisaw in lower species. The function of the
human relaxin-1 peptide is currently unknown, although a
synthetic human relaxin-1 peptide based on the RLN1 se-
quence activates the relaxin receptor (46, 514). The human
RLN1 and RLN2 genes are located sequentially on chro-
mosome 9 at 9p24 adjacent to the INSL4 and INSL6 genes.
In contrast to the cloning of the relaxin genes, the other
members of the relaxin peptide family were discovered as
novel cDNA clones in differential cloning projects or by
screening of EST or genomic databases. All of the new
members retain the relaxin-like preprohormone peptide
structure and a similar gene structure to RLN with two
exons interrupted by an intron situated in the middle of the
C-peptide sequence. INSL3 was discovered by researchers
searching for genes highly expressed in the pig (2) and
1
In this review the terminology species relaxin-1, relaxin-2, relax-
in-3, etc. is used for the relaxin peptides.
Table 1.—Continued
RXFP1 RXFP2 RXFP3 RXFP4
Tissue/Species Rat Mouse Human Rat Mouse Human Rat Mouse Human Human
THP-1
monocytes mRNA
13
mRNA
13
Protein
26
Bone marrow mRNA
2
Skin mRNA
30
mRNA
2
1
Northern blot (237);
2
RT-PCR, immunohistochemistry (239);
3
lacZ reporter expression (293);
4
immuno-histochemistry (255);
5,6,7
RT-PCR (290)
(298) (444);
8
RT-PCR, immunohistochemistry (325);
9,10,11
RT-PCR (443) (354) (406);
12
Northern blot (301);
13,14,15,16
RT-PCR (383) (477) (444)
(228);
17
Northern blot, in situ hybridization (276);
18
receptor autoradiography (67);
19
RT-PCR, immunohistochemistry (322);
20,21
receptor autora-
diography (403) (515);
22
immunohistochemistry (300);
23,24
receptor autoradiography (573) (401);
25
immunohistochemistry (288);
26
receptor
binding (411);
27,28,29
in situ hybridization, receptor autoradiography (333) (416) (189);
30
RT-PCR, Northern blot analysis (455);
31–33
in situ
hybridization, receptor autoradiography (469) (460);
34
RT-PCR (552);
35
receptor autoradiography (519);
36
mRNA by RT-PCR (314);
37
mRNA by
RT-PCR (313);
38
qRT-PCR, Western blotting, immunohistochemistry (550);
39
mRNA by RT-PCR (352);
40
receptor autoradiography (312);
41
in situ
hybridization, receptor autoradiography (508);
42
in situ hybridization, receptor autoradiography (329);
43
in situ hybridization (62);
44
in situ hybrid-
ization, receptor autoradiography (507);
45
in situ hybridization (306);
46
qRT-PCR, Western blotting, immunohistochemistry (551);
47
RT-PCR (259);
48
RT-PCR (286);
49
qRT-PCR (372);
50
qRT-PCR (89);
51
RT-PCR, immunohistochemistry (167);
52
in situ hybridization, receptor autoradiography (486);
53
in situ hybridization, receptor autoradiography (487);
54
qRT-PCR (342).
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
413Physiol Rev VOL 93 JANUARY 2013 www.prv.org
mouse testis (424), and it was initially called Leydig insulin-
like peptide (Ley-I-L) due to its high expression in the tes-
ticular Leydig cells. Subsequent cloning of the INSL3 gene
from numerous mammalian species has confirmed its high
expression in the Leydig cells of mammals (259). The hu-
man INSL3 gene is located on chromosome 19 at 19p13.2.
INSL4 was independently identified as a highly expressed
gene in the human placenta by two groups and was initially
called placenta insulin-like peptide (EPIL) (94) or placentin
(289). The INSL4 gene, like the RLN1 gene, is only present
in higher primates, although there is evidence to suggest
that INSL4 emerged prior to RLN1 (56).
The advent of the EST databases provided an opportunity
to search for insulin-like genes using the conserved A- and
B-chain core sequences. Human INSL5 was identified from
a colon EST library (100), whereas an independent group
discovered mouse INSL5 from a mouse colon EST clone
(236). The mouse clone was originally called relaxin/insu-
lin-like factor 2 (RIF2) and was discovered together with
mouse INSL6 that was called RIF1. Two other independent
groups also discovered mouse (275) and human (317)
INSL6 in testis cDNA libraries. Both INSL5 and INSL6
have been subsequently cloned from numerous mammalian
species and shown to be highly expressed in the colon and
testis, respectively (42). The human INSL5 gene is located
on chromosome 1 at 1p31.3.
The human and mouse RLN3 genes were discovered by
searching the human genome sequence using the relaxin B-
chain (46). The human RLN3 gene is located on chromosome
19 at 19p13.2, 3 mB from the INSL3 gene. The RLN3 gene
has subsequently been identified in all of the sequenced mam-
malian genomes and all other vertebrates including fish, birds,
and amphibians (49, 241, 562). Indeed, bioinformatic analy-
ses showed that RLN3 emerged prior to the divergence of fish,
as RLN3 orthologs are present in Takifugu rubripes and
Danio rerio (562). Phylogenetic analyses also suggest that te-
leosts possess two orthologs that group closely with mamma-
lian relaxin-3, but not with the main circulating form of re-
laxin (410). The high homology between orthologs in the ma-
ture peptide region demonstrates that RLN3 is under strong
purifying selection, unlike the gene encoding relaxin (563).
This is thought to reflect the differences between relaxin and
relaxin-3 receptor interactions and function. Together, the
highly conserved sequence homology and expression of relax-
in-3 in the brain indicate that it is an important neuropeptide
(see sect. IVC).
The RLN3 gene encoding human relaxin-3 is likely to be
the ancestral gene from which a series of gene duplications
gave rise to other members of the relaxin family (562). It is
not possible to assign any order to the divergence between
insulin and the relaxin family peptides; however, insulin,
IGF-I, and IGF-II clearly form a separate subgroup. An
analysis based on peptides from 15 species showed that
human relaxin-3 and INSL5 are closely related to each
other and cluster within a subtree that is distinct from the
group comprising human relaxin-1, human relaxin-2,
INSL3, INSL4, and INSL6 (562). The analysis shown in
FIGURE 1Bbased on human peptide sequences also demon-
strates the close relationship between human relaxin-3 and
INSL5; however, these peptides are clustered within a sub-
tree that also contains human relaxin-1, human relaxin-2,
and INSL6. From this analysis, INSL3 is not more closely
related to human relaxin-2 than the human relaxin-3/
INSL5 peptides. The previous analysis is based on a much
larger data set and provides a comprehensive view of re-
laxin family peptide evolution across diverse species (562).
The discrepancy with the tree based only on the human
peptides reflects the short input sequences available for
comparison, and the fact that the level of homology be-
tween key members of the relaxin peptide family is almost
the same. BLASTP searches (National Center for Biotech-
nology Information) provide measures of amino acid iden-
tity as well as similarity based on conservative changes be-
tween residues that are structurally conserved. The phylo-
genetic tree shown here (FIGURE 1) reflects the fact that human
relaxin-2 is slightly more homologous to human relaxin-3 than
to INSL3. Compared with the core 23-amino acid human
relaxin-2 B chain sequence, human relaxin-3 has 12 identical
residues and an additional 2 conservative changes, while
INSL3 has 10 identical residues and 2 further conservative
changes. Compared with the core 20-amino acid A chain se-
quence of human relaxin-2, INSL3 has 8 identical residues and
no additional conservative changes, and human relaxin-3 has
7 identical residues. The combined B/A chain CLUSTALW
analysis places human relaxin-2 closer to human relaxin-3
than INSL3; however, this is due to an overall difference in
homology of one amino acid. This analysis is not improved by
including the C peptide regions, as these show much greater
divergence in both length and sequence than the A and B
chains. Based on the CLUSTALW and BLASTP analyses, it is
reasonable to conclude that human relaxin-1, human relax-
in-2, human relaxin-3, INSL5, INSL6, and INSL3 all form a
subtree with at least 42% identity and 47% similarity between
members, and that not all branch positions within this group-
ing are likely to be significant.
B. Structure of Relaxins and Insulin-like
Peptides
As mentioned previously (see sect. IIA), the conserved prepro-
hormone structure of the relaxin family peptides (consisting of
a signal peptide, -chain, C-chain, and COOH-terminal A-
chain) necessitates a degree of processing to produce the ma-
ture, active peptide. It is thought that the signal peptide is
removed, followed by the enzymatic removal of the C-peptide,
to form a mature heterodimeric peptide with two disulfide
bonds between the A- and B-chains and an additional intra-A-
chain bond. However, the cleavage of the C-peptide has only
been directly demonstrated for some of the peptides.
BATHGATE ET AL.
414 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
The native relaxin peptide has been isolated and characterized
from numerous species, which allowed the confirmation of a
two-chain, three-disulfide bonded structure similar to insulin
(FIGURE 1A)(42). A similar native structure, generated by en-
zymatic removal of the connecting C-peptide, has been con-
firmed for bovine INSL3 (76) and porcine relaxin-3 (314).
Although there is also some evidence for cleavage of the C-
peptide during INSL6 processing (323), a mature, native pep-
tide has yet to be isolated for both INSL4 and INSL5. Struc-
tural studies on recombinant human relaxin-2 (144) and syn-
thetic INSL3 (435), human relaxin-3 (434), and INSL5 (211)
have enabled the determination of the three-dimensional
structures of the peptides (FIGURE 1C).
The crystal structure of the human relaxin-2 peptide con-
firmed that the overall fold is similar to insulin (144). Further-
more, human relaxin-2 forms an asymmetrical dimer in a
manner resembling insulin, although the human relaxin-2
dimer occurs in a different orientation. Indeed, all relaxin fam-
ily peptides share a common overall fold but have differences
around their termini (FIGURE 1C)(433). Nuclear magnetic
resonance (NMR) studies of human relaxin-3 show that it has
a broadly similar structure to human relaxin-2 and insulin.
However, when compared with the X-ray crystal structure of
human relaxin-2, human relaxin-3 has a more hydrophobic
core with a condensed B-chain
-helix. This “tightness” al-
lows the B-chain tryptophan at position 27 to interact with the
hydrophobic core of the peptide (434). In contrast, the B-chain
-helix in human relaxin-2 is one helix turn longer, forcing the
tryptophan to face away from the core of the molecule where
it is solvent exposed (FIGURE 1C)(434). This orientation of the
COOH terminus of the B-chain of human relaxin-2 is similar
to that of the COOH terminus of INSL5 determined by NMR
(212). In contrast, the NMR structure of INSL3 reveals a
COOH-terminal orientation that is more similar to human
relaxin-3, with the COOH terminus contacting the hydropho-
bic core (434). Despite this, the COOH termini of human
relaxin-3 and INSL3 still show disparity in the precise orien-
tation of the COOH terminus of the B-chain, that has direct
ramifications for the key role of the COOH-terminal trypto-
phan in the activity of these peptides (see sect. IIC). Interest-
ingly, a synthetic INSL4 peptide based on a putative heterodi-
meric mature peptide structure is largely unstructured in solu-
tion (79, 310), bringing into question whether INSL4 is
actually produced as a prohormone in vivo.
C. Structure-Activity Relationships of
Relaxin Family Peptides
1. Structural features of relaxin necessary for
biological activity
Early studies on relaxin highlighted that both A- and B-
chains are necessary for activity since treatment of relaxin
with reducing agents destroyed its biological activity (174).
The initial effort to identify relaxin peptides in a range of
different species was the first step in determining the struc-
ture-activity relationships of relaxin family peptides, their
receptors, and downstream signaling effectors. Alignment
of the amino acid sequences of relaxin peptides from rat,
tiger shark, dogfish, rabbit, horse, skate, minke whale, dog,
human, dolphin, and cow allowed the identification of key
conserved residues, i.e., those most likely to be important
for the structure-activity relationships of relaxin (470).
Aside from the essential cysteine residues, these alignments
highlighted a conserved RxxRxxI/V motif located in the
B-chain
-helix (FIGURE 5). The X-ray crystal structure of
relaxin indicates that the arginine residues are located on
the first and second loop of the
-helix and face away from
the core of the molecule (FIGURE 1C)(144). The outward-
facing arginines form the receptor binding site for relaxin,
in conjunction with an isoleucine or valine residue that is
located on the same face of the B-chain
-helix (452, 513).
Substitution of the arginine, isoleucine, or valine residues in
this motif markedly reduces or obliterates activity (FIGURE
5) (452, 513), highlighting that these residues are critical for
the activity of the relaxin peptide.
Analysis of the relaxin family peptides showed that the
A-chain of the relaxins had greater sequence divergence
than the B-chains (470) (FIGURE 1A). Other than the con-
served cysteine residues, only G14 seems to be highly con-
served (FIGURE 5) and is essential for maintaining chain
flexibility and structure (73). Therefore, it was assumed that
the relaxin peptide A-chains would play a less important
role in receptor binding and activation than the B-chain.
Truncation of the A-chain of human relaxin-2 results in a
peptide that progressively loses the ability to bind to and
activate both RXFP1 and RXFP2 (232). It was thought that
this might be due to the loss of structure in the B-chain
resulting from the truncation. However, a recent study
demonstrated that targeted point-mutations within the hu-
man relaxin-2 A-chain can selectively alter receptor binding
and activation profile: alanine substitution at the A16 and
A17 positions, (T16A and K17A, respectively) did not
markedly alter the cAMP response to RXFP1 activation,
but enhanced cAMP accumulation following peptide bind-
ing to RXFP2 compared with wild-type control (FIGURE 5);
furthermore, alanine substitution of residues A22 and A23
(R22A and F23A, respectively) markedly reduced the human
relaxin-2 activity at RXFP2 without reducing binding or ac-
tivity at RXFP1 (FIGURE 5) (410). Additionally, double ala-
nine substitutions at positions A19 and A20 (S19A and L20A,
respectively) also reduced peptide binding and activity at
RXFP1, again suggesting that residues of the human relaxin-2
A-chain are important for receptor binding and activation
(FIGURE 5) (73). This study also highlights differences in the
mechanisms by which relaxin can bind to and activate RXFP1
compared with RXFP2 (FIGURE 1C). Differences in the mech-
anism of binding and activation between RXFP1 and RXFP2
are further supported by the species-specific nature of these
structure-activity relationships; thus neither mouse nor rat re-
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
415Physiol Rev VOL 93 JANUARY 2013 www.prv.org
laxin can bind to or activate RXFP2 (454), and human relax-
in-3 has only a poor affinity for RXFP2 (45).
2. Relaxin-3: structure and function
While relaxin-3 requires both A- and B-chains to bind to
and activate RXFP1, it only requires the B-chain to bind
and activate RXFP3 and RXFP4 (314) as the B-chain of
relaxin-3 is a low-affinity agonist at both receptors (313,
314). Replacement of the human relaxin-3 A-chain with the
INSL5 A-chain (R3/I5) does not influence RXFP3 binding
or activation, but the chimeric peptide displays markedly
reduced ability to bind to and activate RXFP1 (see sect. IIC)
(312). Furthermore, truncation of the human relaxin-3 A-
chain produces a peptide that no longer binds to or activates
RXFP1 but retains full activity at RXFP3 (232).
Mutagenesis of key human relaxin-3 B-chain residues suggests
that R8, R16, I15, and F20 are important for human relaxin-3
binding to both RXFP3 and RXFP4 (FIGURE 5) (299), whereas
R12 is only important for human relaxin-3 binding to RXFP3
(FIGURE 5) but not to RXFP4 (299). Additionally, the muta-
tion of R26A and W27A causes a loss of inhibition of forsko-
lin-stimulated CRE activation by the peptide (299), suggesting
that these residues are essential for RXFP3-mediated inhibi-
tion of cAMP (FIGURE 5). These mutagenesis studies facili-
tated the development of a high-affinity RXFP3 antagonist
(R3/I5) that has a truncated human relaxin-3 B-chain (re-
Human relaxin-2 A-chain
T16A or K17A enhanced cAMP accumulation
at RXFP2 but not RXFP1
S19A/L20A reduced binding and activity at
RXFP1
R22A or F23A markedly reduced activity at
RXFP2 without reducing binding or activity
at RXFP1
Human relaxin-2 B-chain
R13, R17 and I20 involved in binding to
RXFP1
Truncation of B-chain N-terminus results in
progressive loss of binding to RXFP1 and
RXFP2. Truncation of C-terminus past G24
and W27 results in loss of binding for RXFP1
and RXFP2 respectively
Human relaxin-3
A-chain
Truncation of A-chain
N-terminus-progressive
loss of binding to RXFP1
but RXFP3 binding
not affected until
truncation > R8.
Human relaxin-3
B-chain
R8, I15, R16 and F20
involved in binding to
RXFP3 and RXFP4
R12 involved in binding
to RXFP3
R26 and W27 involved in
activation but not binding
to RXFP3 and RXFP4
INSL3 A-chain
C10S/C15S or C10del/C15del devoid of cAMP
signalling activity but retained affinity for RXFP2
N-terminus truncation (AAATNPA; AAATNPAR or
AAATNPARY) devoid of cAMPsignallingbut retained
binding to RXFP2
INSL3 B-chain
W27critical for RXFP2 binding and activation
H12, R16, V19 and R20 all contribute to binding
together with W27
N-terminus truncation (PTPEMREK) - loss of cAMP
signallingbut retained binding affinity at RXFP2
FIGURE 5. Structure-activity relationships for the relaxin family peptides human relaxin-2, human relaxin-3,
and insulin-like peptide 3 (INSL3). The three peptides share common structural features including the two
disulfide bonds between the A and B chains and the intrachain disulfide bond in the A chain. In spite of these
similarities, all of the peptides bind to their cognate receptors in characteristic ways. Binding of human
relaxin-2 to RXFP1 involves the RxxxRxxI/V motif in the B chain (144, 470), and mutations in the A chain can
alter activity at both RXFP1 and RXFP2 (232, 410); binding of INSL3 to RXFP2 involves a similar motif but
displaced one turn along the
-helix of the B chain (74–76), whereas NH
2
-terminal deletions alter functional
responses while retaining binding (74). The human relaxin-3 B chain contains motifs associated with both
binding and activity (299, 434), and the A chain can be shortened without affecting activity (232). Similar
motifs that occur across peptides probably account for the cross-reactivity seen for some peptides at other
RXFP receptors.
BATHGATE ET AL.
416 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
moving R26 and W27), a G23A mutation, and the A-chain of
INSL5 (299) to produce a peptide that does not bind to
RXFP1. More recently, a single-chain antagonist has been de-
veloped that utilizes a modified B-chain of this peptide. This
has the -chain of R3/I5 with the cysteine residues mutated
to serine (C10S-C22S) (213).
3. INSL3: structural features associated with
specific interaction with RXFP2
INSL3 interacts with RXFP1 only at extremely high concen-
trations (199, 239), suggesting that it is unlikely to bind to or
activate RXFP1 in a physiological setting. Mutagenesis studies
on synthetic INSL3 peptides have demonstrated that INSL3
utilizes numerous amino acids in the B-chain to bind to and
activate RXFP2. W27 of the INSL3 B-chain is critical for
RXFP2 binding and activation (74, 77), although other single
amino acid replacements with either alanine or valine
throughout the INSL3 B-chain suggested that there were also
other important residues for receptor binding in addition to
W27 (FIGURE 5). However, while these individual substitu-
tions only slightly reduced the binding of INSL3 to RXFP2, in
combination they produce a large effect (75, 435). Thus the
combined mutation of H12A, R16A, V19A, R20A, and
W27A results in a mutant that cannot bind to RXFP2, high-
lighting the necessary contribution of all of these residues (FIG-
URE 5) (435). There have also been a number of INSL3 muta-
tions that have been identified in both normal and patient
populations (see sect. VIIB1).
Studies on INSL3 B-chain analogs have suggested that while
the B-chain alone can still bind to RXFP2, these single chain
peptides cannot activate the receptor and therefore act as
functional antagonists (121, 464). A similar effect is pro-
duced within the peptide dimer by truncation of the INSL3
A-chain: while deletion of up to six residues has no effect on
the full agonist activity of the peptide, deletion of up to 10
residues from the NH
2
terminus of the INSL3 A-chain pro-
duces a peptide that still binds to RXFP2, but no longer
increases cAMP accumulation (74). These derivatives are
specific competitive inhibitors of INSL3 at RXFP2 (74).
Similarly, as with the native INSL3 A-chain, deletion of
eight NH
2
-terminal residues of the B-chain produces an
INSL3 peptide that retains binding affinity for RXFP2, but
lacks the cAMP signaling function (74). Replacing any of
these eight B-chain residues individually with alanine does
not affect receptor signaling, nor does replacing all eight
residues of the A-chain with alanine (74). Alteration of the
cysteine residues of the intra-chain A-chain disulfide bond
(C10S/C15S; or C10del/C15del) also produced peptides
that retain RXFP2 binding but do not activate cAMP sig-
naling (FIGURE 5) (587). All of these studies highlight that
while both the A- and B-chains are necessary for INSL3
activation of RXFP2, the mechanism of action is different
from the relaxin-RXFP1 interaction (232).
III. MOLECULAR BIOLOGY, STRUCTURAL
FEATURES, AND FUNCTIONAL
DOMAINS OF RXFP RECEPTORS
The relaxin family peptide receptors are four highly con-
served family A GPCRs that fall into two groups based on
their architecture and signaling properties. The RXFP1 and
RXFP2 genes have introns, and there are a large number of
splice variants many of which result in proteins whose phys-
iological function is yet to be established. In contrast, the
genes encoding RXFP3 and RXFP4 are intronless. Both
RXFP1 and RXFP2 have a large extracellular NH
2
termi-
nus containing a leucine-rich repeat (LRR) domain that is
responsible for high-affinity ligand binding (this is comple-
mented by a low-affinity binding interaction within ex-
oloop 2) and, uniquely, an LDLa module that is essential for
signaling and influences receptor translocation. While
RXFP1 and RXFP2 show similarities in ligand binding and
signaling, only RXFP1 has a COOH terminus containing
motifs that induce the formation of highly sensitive protein
signaling complexes termed signalosomes. The third intra-
cellular loop is essential for G protein coupling in all RXFP
receptors. RXFP1 and RXFP2 couple to G
s
and to G
OB
which modulates this effect, but only RXFP1 is able to
couple to G
i3
involving R752 in the COOH-terminal tail
to produce a delayed surge in cAMP accumulation. On the
other hand, RXFP3 and RXFP4 couple solely to G
i
and
G
O
proteins with the pattern of coupling dependent on the
cell type in which the receptor is expressed.
A. Molecular Biology of Relaxin Family
Peptide Receptors
1. Leucine-rich repeat containing receptors RXFP1
and RXFP2
RXFP1 and RXFP2 are family A G protein-coupled receptors
(GPCRs), similar to the rhodopsin receptor, that specifically
belong to the leucine-rich repeat-containing GPCR (LGR)
subfamily (reviewed in Ref. 18). Studies of LGRs from differ-
ent species suggest that the three subtypes of LGRs (A, B, and
C) originated during the early evolution of metazoans (241).
Generally, LGRs contain a leucine-rich repeat (LRR) domain,
a hinge region, seven transmembrane-spanning domains, and
a COOH-terminal tail. Type A LGRs include the FSH recep-
tor, the LH receptor, and the TSH receptor, each of which
contains nine LRRs before the hinge region. The type B LGRs
consist of three members: LGR4, LGR5 and LGR6. Both
LGR4 and LGR5 have 17 LRRs, whereas LGR6 has 13 LRRs.
All the receptors from this subgroup have a different hinge
region to the type A LGRs. The type C LGRs, RXFP1 (LGR7)
and RXFP2 (LGR8), are differentiated by the presence of a
LDLa module at the extreme NH
2
terminus, leading into 10
LRRs and a unique hinge region prior to the transmembrane-
spanning domain (FIGURE 3).
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
417Physiol Rev VOL 93 JANUARY 2013 www.prv.org
There is a high degree of conservation within the sequence
of RXFP1 and RXFP2 from fish to mammals (240), with
human and rodent orthologs exhibiting 90% sequence
similarity (238, 239, 455). Human RXFP1 is located on
chromosome 4q32.1, and RXFP2 is located on chromo-
some 13q13.1; the receptors share 60% amino acid se-
quence identity and 80% homology.
The genes for RXFP1 and RXFP2 are very large, covering
60 kB and both contain 18 exons. Alternative splicing
occurs commonly within LGRs, and there are a number of
splice variants of all type A LGRs (188, 319, 366, 377,
530), as well as LGR4 and LGR6 (375). To date, 29 splice
variants of RXFP1 and RXFP2 have been described result-
ing in multiple potential protein products including frag-
ments consisting only of the transmembrane or the extra-
cellular region (282, 375, 456). A selection of these variants
has been characterized in transfected cell systems, and al-
though some of the transmembrane-containing products
translocate to the plasma membrane, others are retained
inside the cell. Most of the transmembrane containing vari-
ants that are expressed at the cell surface are unable to bind
ligand (375). However, a RXFP2 variant that lacks only the
exon that encodes the LDLa module is able to bind ligand
but does not activate cAMP production, suggesting that this
variant may be a native binding protein that could modulate
INSL3 action in vivo (456). Products encoding only the
extracellular portions of the receptors, including variants
that express the LDLa module alone, can be secreted from
the cell (456). These variants could therefore act as func-
tional RXFP1 antagonists by binding and preventing re-
laxin activation of RXFP1. It is postulated that the native
variants that are expressed in rodent tissues could act as
endogenous antagonists (457). Some of the other splice
variants have also been demonstrated to modulate the func-
tion of the wild-type receptors. Three RXFP1 splice variants
occur in human fetal membranes, and in vitro studies in
transfected cells suggest that these variants can form het-
erodimers with RXFP1 and consequently modulate the
function of the receptor (282).
Construction of phylogenetic trees to examine the evolu-
tionary relationship between human RXFP receptors and
other family A GPCRs confirms that RXFP1 and RXFP2
are closely related to one another, with 59% sequence iden-
tity and 80% similarity based on conservative amino acid
changes, and also to the three glycoprotein hormone recep-
tors for thyroid stimulating hormone (TSHR), luteinizing
hormone/chorionic gonadotropin (LHCGR), and follicle
stimulating hormone (FSHR). RXFP1 and RXFP2 are
somewhat less related to the LRR-containing receptors
LGR4, LGR5, and LGR6 (22–25% identity, 4647% sim-
ilarity) (FIGURE 2A). LGR4 and LGR5 have recently been
shown to regulate Wnt/
-catenin signaling in response to
R-spondins (90). RXFP1 and RXFP2 are also (somewhat
surprisingly) related to family A bioamine receptors such as
-adrenoceptors (ARs) and adenosine receptors, as well as
additional peptide receptors (FIGURE 2A).
2. Small peptide receptor-like receptors RXFP3 and
RXFP4
RXFP3 and RXFP4 were deorphanized 1 year after RXFP1
and RXFP2 (42, 313–315) and are most closely related to
small peptide receptors, with a relatively short NH
2
-termi-
nal domain common among family A GPCRs (FIGURE 2B).
As a consequence of this short NH
2
-terminal domain,
RXFP3 and RXFP4 are significantly smaller (400 resi-
dues) than RXFP1 and RXFP2 (700 residues). Human
RXFP3 is located on chromosome 5p13.2 and RXFP4 on
chromosome 1q22. Unlike RXFP1 and RXFP2, the genes
do not contain introns and thus have no splice variants.
RXFP3 and RXFP4 are closely related to one another (43%
sequence identity and 60% similarity) and to a subgroup
comprising the angiotensin II type 1 receptor (AGTR1), the
apelin receptor (APLNR) and the chemokine receptor
CCR1, and to a second subgroup containing neuropeptide
receptors including members of the somatostatin and opi-
oid receptor families and the neuropeptide B/W receptor
(NPBWR1) (FIGURE 2B). They are also more distantly re-
lated to various chemokine receptors (CXCR7, CXCR4,
CCR1, and CCR9) and to the formyl peptide receptors
FPR2 and FPR3 (FIGURE 2B). However, in spite of the close
similarity between the cognate ligands for these receptors,
the RXFP1/RXFP2 and RXFP3/RXFP4 receptor subgroups
are not highly related to each other (FIGURE 2C), and in fact,
the only significant homology between RXFP1 and RXFP3
is within the highly conserved TM7 motif NSxLNP(I/V)Y
that is essential for G protein coupling as well as neighbor-
ing key residues in helix 8. RXFP1 and RXFP4 show
slightly higher homology, although again the regions of
similarity encompass only short motifs in TM3, TM6, and
TM7. Instead, both receptor subgroups are related to other
family A GPCRs such as the
2
-AR and the adenosine A2B
receptor. In essence, these bioamine receptors occupy a
bridge between the distinct RXFP1/RXFP2 and RXFP3/
RXFP4 subgroups (FIGURE 2C). For example, the
2
-AR has
homologies with each of the RXFP1 and RXFP3 receptors,
with 23% identity/39% similarity to RXFP1 and 21% iden-
tity/38% similarity to RXFP3.
B. Structural Features and Mode of Ligand
Activation of the LRR-Containing
Receptors RXFP1 and RXFP2
1. General features of RXFP1 and RXFP2
The high degree of similarity between RXFP1 and RXFP2
has facilitated the identification of the functional domains
of the two receptors. Initial work used relaxin-3 because of
its selectivity for RXFP1 over RXFP2, and utilizing chime-
BATHGATE ET AL.
418 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
ric receptors revealed that the peptide needs to interact with
both the ectodomain and exoloop 2 of the transmembrane
domain of RXFP1 in order for the full binding and cAMP
signaling profile to be achieved (501). The bimodal interac-
tion of relaxin with RXFP1 was later shown to be a feature
of both receptors that both contain two ligand-binding
sites: a high-affinity site within the ectodomain and a lower
affinity site within the transmembrane region (FIGURE 3)
(199). Evidence for two distinct binding sites was also pres-
ent in functional assays, with the high-affinity ectodomain
site inducing cAMP accumulation with higher efficiency
than the lower affinity transmembrane site (TABLE 2) (199).
Studies with chimeric and truncated receptors showed that
both sites are necessary for optimal binding and signal
transduction (199). In addition to the LRR-region of the ect-
odomain, the NH
2
-terminal LDLa module is essential for sig-
naling, although to date there is no evidence for a role of this
domain in ligand binding (TABLE 2) (231, 280, 456). Thus
RXFP1 and RXFP2 have two ligand binding sites, and recep-
tor binding followed by activation is currently thought to re-
quire the LRR, extracellular loops, and the LDLa module.
In a manner similar to many other class A GPCRs, recent
BRET studies have also shown that both receptors form
homo- and heterodimers (282, 510, 511). For RXFP1 and
RXFP2, dimerization occurred in the absence, and was in-
dependent, of ligand occupation of the receptor (510, 511).
Furthermore, dimerization between the haloreceptor and a
number of splice variants (encoding only the LDLa module,
and up to 8 LRR) was evident, and dimers were present at
all stages of receptor translocation from the endoplasmic
reticulum to the plasma membrane (282), suggesting an
important role of dimerization over the lifetime of the re-
ceptor. Dimerization was also evident between the halore-
ceptor and a transmembrane-only domain receptor, al-
though the BRET
2
ratio was decreased compared with that
obtained for haloreceptor dimers (510, 511). This suggests
that while the transmembrane domain is sufficient for
dimerization, the ectodomains play an important role in
stabilizing the oligomer.
A functional consequence of dimerization is that both re-
ceptors display negative cooperativity (510, 511). In con-
trast to allosterism, where binding of a molecule to a single
site influences the properties of a second site, the two bind-
ing sites do not have a fixed affinity in proteins exhibiting
negative cooperativity. Instead, the affinity of each remain-
ing unoccupied receptor binding site decreases as occu-
pancy increases. There are two intriguing, functionally rel-
evant consequences of negative cooperativity at RXFP1 and
RXFP2: 1) this phenomenon increases the functional range
of the ligand over a bigger concentration range, and 2) the
ligand residence time at the receptor shortens as the free
ligand concentration increases, potentially allowing selec-
tive activation of different signaling pathways (476). The
negative cooperativity concentration-response curve for
RXFP2 was bell-shaped, suggesting that binding occurs in a
trans manner within the dimer: at low concentrations,
INSL3 binds to the high-affinity site on receptor 1 and the
low-affinity site on receptor 2, with partial ligand dissocia-
tion and increased free ligand concentrations; another
INSL3 molecule binds in the same manner to the remaining
two sites, accelerating the dissociation of the first INSL3
molecule; and finally, at very high concentration of free
ligand, this effect is lost due to two additional INSL3 mol-
ecules binding to the free high- and low-affinity sites, thus
preventing further cross-linking (510). Contrastingly, the
negative cooperativity concentration-response curve for re-
laxin binding to RXFP1 was linear, although the absence of
a protein structure precludes conclusions regarding the
functional consequences of this observation (511)].
2. A unique feature of RXFP1 and RXFP2: the LDLa
module
The LDLa module was originally defined by the structure of
the LDL receptor (110, 571) but has since been identified in
a number of other proteins including the very-low-density
lipoprotein receptor (177), the LDL receptor-related pro-
tein (218), renal glycoprotein gp330 (223, 326), the C9
component of complement (123, 494), the Tva receptor for
Rous sarcoma virus (39), and retina- and brain-specific neu-
ropilin, and tolloid-like protein 1 transmembrane protein
(500). RXFP1 and RXFP2 are the only known human GPCRs
to contain this module (42). The NMR solution structure of
the RXFP1 LDLa module was recently solved (FIGURE 3) and
shows the typical fold generated by the essential six cysteine
residues, and the anticipated incorporation of a calcium ion by
a largely conserved motif of acidic residues (231).
Mutagenesis studies of the LDLa region have highlighted the
essential role of this module in receptor signaling. Mutation of
conserved residues, those that compromised correct folding,
or indeed entire deletion of the LDLa module of either RXFP1
or RXFP2 resulted in receptors that when occupied are unable
to increase cAMP, despite these mutant receptors maintaining
intact ligand binding profiles (231, 280, 456). Mutation of
two conserved cysteines (C47A and C53A), or mutation of a
residue within the calcium-binding motif (D58E), resulted in
mutant RXFP1 receptors that were unable to increase
cAMP accumulation in response to either human relaxin-1
or human relaxin-2 (280). Other mutations that affect the
ability of the LDLa module to correctly fold (C5S and
C18S, which disrupts the first disulphide bond) also results
in a receptor that is unable to increase cAMP accumulation
(TABLE 2) (231). Additionally, when the RXFP1 LDLa mod-
ule is swapped for the LDLa module contained within the
second ligand-binding domain of the LDL receptor, the chi-
meric receptor was unable to increase cAMP accumulation
in response to human relaxin-2 but bound ligand normally
(231). These data together with further mutagenesis of spe-
cific residues in the RXFP1 LDLa module indicate that this
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
419Physiol Rev VOL 93 JANUARY 2013 www.prv.org
domain mediates signaling by a specific side-chain-driven
interaction (231).
The LDLa module may also play a role in receptor matura-
tion and delivery to the cell surface. A high proportion of
full-length RXFP1 receptors were found to exist in an im-
mature form containing high mannose-type N-linked oligo-
saccharides, consistent with retention of the receptor within
the endoplasmic reticulum (280). This has previously been
reported for other glycoprotein hormone receptors includ-
AKAP79 β-arrestin
RXFP1 + + + + + + + ++
LDLa- + + - - - - - NT NT
LRR + - - - - - - NT NT
TM - + - - - - - NT NT
t703 + + + - + + - NT NT
t747 + + + - + + - NT NT
R752A +++ -++-
NT NT
S755A ++++++-
NT NT
S704A + + + + + + + +-
D637A + + + + + NT NT NT NT
C47A
C53A
++ - --NTNT
NT NT
D58E ++ - --NTNT
NT NT
C5S
C18S
++ - --NTNT
NT NT
D231N E233Q
E277Q D279N -NTNTNTNTNTNT
NT NT
--+NTNTNTNT
NT NT
Gαi3 GαoB Gαs
NFκB
cAMP
CRE
Binding
ECL
Binding
LRR
IC3 Fragments
615-629, 619-
629K Palmitate
TABLE 2. Functional domains of the human RXFP1 receptor determined using receptor fragments or
site-directed mutagenesis. The effect of deletion of regions or site-directed mutagenesis was determined on
labeled relaxin binding to the leucine-rich repeat region (LRR), extracellular loops (ECL); on cAMP generation
determined by activation of CRE or NF
B reporter genes; on signaling mediated by G
s
,G
OB
,orG
i3
; and on
binding to AKAP79 or
-arrestin. The modified RXFP1 receptors utilized included RXFP1 lacking the low-density
lipoprotein receptor type A module (LDLa), consisting of the NH
2
-terminal domain of RXFP1 (LRR), the
transmembrane spanning region (TM), truncated RXFP1 receptors (t703 and t747), receptor variations
generated by site-directed mutagenesis, and fragments of intracellular loops of RXFP1. , Response or
interaction present; , no response or interaction; , reduced response or interaction. NT, not tested. Data
are from References 199, 204, 231, 280, 456, 475, and 501.
BATHGATE ET AL.
420 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
ing the LH receptor (18, 419, 520) and the FSH receptor
(115, 428) and is suggested to be a form of control over cell
surface expression. Interestingly, RXFP1 lacking an LDLa
module was expressed as the mature form, as was a chime-
ric RXFP1 receptor containing the LDLa module of RXFP2
(280). Furthermore, mutation of a conserved glycosylation
site within the LDLa module results in a receptor with a
reduced ability to generate cAMP, which was attributable
to a decrease in cell surface expression (280, 572). In
RXFP2, mutations of amino acid residues that form the
disulfide bond or coordinate Ca
2
binding in the LDLa
module (C71Y and D70Y) also reduced cell surface expres-
sion (64). Taken together, this suggests that the LDLa mod-
ule may impair, or negatively regulate, receptor expression
at the plasma membrane.
3. The LRR region and its importance for receptor
translocation, ligand binding, and signal
transduction
Glycosylation is a posttranslational modification common
to many GPCRs and is important for receptor delivery to
the cell surface, ligand binding, and signal transduction. In
addition to the glycosylation sites within the LDLa module,
RXFP1 contains a number of glycosylation sites within the
LRRs (asparagines at N105, N250, N303, and N346) that
are all important for both cell-surface localization of the
receptor and full cAMP signaling efficacy, but have no role
in ligand binding (572). As yet, the glycosylation status of
RXFP2 has not been studied in detail.
Both RXFP1 and RXFP2 contain two peptide-binding sites:
a high-affinity site in the LRR region and a lower-affinity
site in the extracellular loop region (TABLE 2) (199, 501).
Modeling of the LRR region of RXFP1 based on the crystal
structure of the porcine ribonuclease inhibitor (a protein
with LRRs) allowed in silico peptide docking and targeted
receptor mutagenesis (78). These studies revealed that the
site most likely to bind relaxin occurs at a 45° angle across
five of the parallel LRRs (78). Relaxin binding to these
LRRs occurs through synchronized chelation of two argi-
nines in the relaxin B-chain (R13 and R17: part of the
relaxin binding motif), neutralization of charge from acidic
groups within the concave face of the LRRs (E277 and
D279 in LRR8, and D231 and E233 in LRR6), and gener-
ation of hydrogen bonding (78). This is stabilized by a hy-
drophobic interaction between the relaxin B-chain isoleu-
cine (I20) and a receptor-based cluster of W180 and I182
within LRR4, and L204 and V206 within LRR5 (78). De-
letion of any of these residues within the receptor abolished
relaxin binding (78). Although these residues are also pres-
ent on RXFP2, evidence suggests that human relaxin-2
binds to RXFP2 in an INSL3-like manner (458).
As outlined in section IIC, INSL3 utilizes different residues
in the B-chain to interact with RXFP2 compared with the
relaxin-RXFP1 interaction (TABLE 3). However, human re-
laxin-2 contains most of these residues and can mimic the
INSL3 interaction. By using the NMR solution structure of
INSL3, and molecular modeling of the RXFP2 LRR based
on the crystal structure of the Nogo receptor (47% amino
acid sequence homology), a recent study identifies the
RXFP2 residues that directly interact with the INSL3 B-
chain (459). With the use of targeted mutagenesis of both
peptide and receptor, in conjunction with in silico docking
of the INSL3 B-chain to the LRR of RXFP2, seven residues
within RXFP2 were identified. Thus within the B-chain of
INSL3, residue R16 is predicted to interact with RXFP2
D227, INSL3 H12 with RXFP2 W177, INSL3 V19 with
RXFP2 I179, INSL3 R20 with RXFP2 E229 and D181, and
INSL3 W27 with RXFP2 F131 and Q133 (459). The pri-
mary determinants of INSL3 binding are the peptide residues
R20 and W27 within the B-chain. Although five of the identi-
fied RXFP2 residues are conserved in RXFP1, the two non-
conserved residues (D181 and F131) interact with essential
residues within the INSL3 binding motif: R20 and W27 (459).
Thus this model potentially provides an explanation for the
lack of INSL3 binding at RXFP1.
4. Evidence for low-affinity relaxin family peptide
binding to extracellular loops in RXFP1 and
RXFP2
In addition to the relatively well-defined high-affinity ligand-
binding site within the LRRs, both RXFP1 and RXFP2 have
an additional lower affinity binding site presumed to be within
the transmembrane exoloops of the receptors. The existence of
this site was first suggested by a study that utilized chimeric
receptors, which exploited the inability of relaxin-3 to either
bind to or activate RXFP2 (501). The binding affinity and
cAMP signaling efficacy of relaxin-3 was reduced at an
RXFP1/2 chimera (RXFP1 ectodomain, fused to the trans-
membrane and COOH terminus of RXFP2) compared with
RXFP1. However, replacement of the second extracellular
loop (ECL2) of RXFP2 with that of RXFP1 restored relaxin-3
binding and signaling, thereby identifying ECL2 as a second-
ary binding site (TABLE 3).
This additional low-affinity site has since been confirmed
for other relaxin peptides at both RXFP1 and RXFP2.
These studies used the same chimeric receptors described
above and exploited the preference of rat relaxin for RXFP1
over RXFP2, and the preference of INSL3 for RXFP2 over
RXFP1 (199). Thus primary binding is directed by the ectodo-
main, with an influence of the low-affinity site within the
transmembrane ECLs. This mechanism is also reflected in
cAMP signaling efficacy profiles; thus the high-affinity site sig-
nals to cAMP accumulation with high efficiency, and the low-
affinity site with low efficiency. The precise location of the sites
has yet to be determined, although it is presumed to be within
the ECLs and may involve an interaction with the A-chain of
the relaxin family peptides (TABLE 3).
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
421Physiol Rev VOL 93 JANUARY 2013 www.prv.org
5. Functional roles of the intracellular loops
The third intracellular loop appears to play an important
functional role in coupling RXFP1 to G
s
and thus is
important in controlling increased cAMP accumulation
in response to receptor activation (see sect. VIA). Peptide
fragments derived from the NH
2
-terminal region of the
third intracellular loop [IC3; residues 615–629, and
619629-Lys(Palm)] increased adenylyl cyclase (AC) ac-
tivity independently of relaxin (475) (TABLE 2). The same
peptides functionally “antagonized” the cAMP response
to relaxin activation of RXFP1 endogenously expressed
in rat striatum and rat cardiac muscle (475). Further-
more, a synthetic peptide derived from the COOH termi-
nus of G
s
(residues 385–394) inhibited both the AC
activity stimulated by relaxin and the increased cAMP
stimulated by the IC3 peptides alone (475), suggesting
that G
s
interacts with the IC3 of RXFP1 (FIGURE 3).
Many other GPCRs interact with G proteins by their
intracellular loops, and there is much evidence to support
this interaction site (143, 217, 284, 431). It is tempting to
speculate that IC3 not only directs coupling to G
s
, but
also directs coupling to G
oB
for both RXFP1 and
RXFP2 (see sect. VIA): truncation of either receptor after
helix 8 did not affect these components of the cAMP
signaling response, suggesting a common interaction site,
although this remains to be demonstrated. Similarly, the
constitutive activity of RXFP1 mediated by G
s
(and
G
␤␥
) activation of adenylyl cyclase 2 (AC2) (see sect.
IIIB6; Ref. 201) is also likely to depend on G
s
coupling
to ICL3.
6. The COOH-terminal tail of RXFP1 controls both
cAMP signaling through the delayed pathway and
signalosome formation
Only RXFP1, but not RXFP2, increases cAMP accumula-
tion by coupling to G
i3
(198). Activation of adenylyl cy-
clase 5 (AC5) to increase cAMP occurs downstream of
G
i3
,byaG
␤␥
-PI3K-PKC
pathway (see sect. VIA). The
coupling of RXFP1 to this pathway involves the final 10
amino acids of the receptor COOH terminus and absolutely
Receptor Relaxin INSL3 Relaxin INSL3
RXFP1 9.75 5.68 9.39 ND
RXFP2 8.96 9.71 7.88 8.27
RXFP1/2 9.48 8.57 9.08 8.5
RXFP2/1 9.18 10.05 8.17 8.82
9.66 7.67 - -
9.37 9.69 - -
9.59 ND 8.91 ND
9.85 ND 9.07 ND
9.28 ND 8.78 ND
Binding - pKD, pKi cAMP - pEC50
RXFP1
/2ECL1
RXFP2-BP
RXFP1-BP
RXFP1
/2ECL2
RXFP1
/2ECL3
N.B. Top 6 rows are means of published values
TABLE 3. Functional domains of RXFP1 and RXFP2 receptors determined using receptor chimeras. RXFP1
residues are outlined in black, with RXFP2 in gray. Binding is determined using radiolabeled relaxin and
expressed as log [concentration]. cAMP accumulation was measured in response to relaxin or INSL3. The
top 6 rows are the means of published data. ND, not detectable. Data are from References 199 and 501.
BATHGATE ET AL.
422 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
requires R752 within these residues (FIGURE 3), as trunca-
tion of the COOH terminus or substitution of R752 selec-
tively removes coupling to the G
i3
pathway (TABLE 2)
(204). As yet, the precise mechanism by which the terminal
10 residues of RXFP1 direct G
i3
coupling is unclear. There
are a number of possibilities including direct G
i3
coupling,
G
i3
coupling induced by receptor phosphorylation, or re-
cruitment of scaffolding proteins for colocalization of the
receptor with G
i3
. However, it would be unprecedented
for G protein coupling to occur directly with the RXFP1
COOH-terminal tail.
The COOH-terminal tail of RXFP1 also controls a small
degree of constitutive activity (201) and only RXFP1, but
not RXFP2, constitutively couples to AC2 by an associ-
ation between helix 8 and A-kinase anchoring protein 79
(AKAP79). This protein complex also facilitates G
s
and
G
␤␥
-mediated stimulation of AC2 to increase cAMP ac-
cumulation in response to subpicomolar concentrations
of relaxin (see sect. VIA). The level of cAMP generated by
this complex is tightly controlled by the activity of pro-
tein kinase A (PKA)-stimulated phosphodiesterase (PDE)
4D3, which is localized to the receptor COOH terminus
by an association between
-arrestin 2 and S704 of
RXFP1. This signalosome, directed and maintained by
the RXFP1 COOH-terminal tail, mediates a cAMP re-
sponse to low concentrations of peptide and may provide
a novel cellular response to low levels of relaxin in some
physiological situations.
The RXFP1 COOH-terminal tail also contains a number of
potential consensus sequences for phosphorylation and
protein-protein interactions and is the region of most vari-
ation between RXFP1 and the highly related RXFP2 (203),
thus representing an area of functional divergence between
the two receptors which may relate to the more varied phys-
iological roles of relaxin in relation to INSL3.
C. The Mode of Ligand Activation of the
Small Peptide Receptor-like Receptors
RXFP3 and RXFP4
As noted previously, the small NH
2
-terminal domains of
RXFP3 and RXFP4 are markedly different from RXFP1
and RXFP2. The precise mode by which relaxin-3 and
INSL5 bind to and activate their receptors is currently
unknown. However, current evidence suggests that the
B-chain of relaxin-3 contains all the RXFP3 contact
points. It is postulated that the residues in the B-chain
central helix R8, R16, I15, and F20 that are important for
human relaxin-3 binding may interact with the RXFP3
ECLs or NH
2
-terminal domain, whereas residues R26 and
W27, which are essential for receptor activation, may inter-
act with the RXFP3 transmembrane helices (291, 594). The
precise residues in both RXFP3 and RXFP4 that interact
with these amino acids are currently unclear, although one
study has extensively examined the potential roles of the
extracellular domains and transmembrane helices in recep-
tor function (TABLE 4) (594). This study examined the
binding and activation of RXFP3 and RXFP4 chimeras
by INSL5 and human relaxin-3 to determine regions of
the receptor necessary for ligand interaction. A chimera
of RXFP3 (transmembrane domain, loops, and COOH-
terminal tail) with the NH
2
-terminal region of RXFP4
(CR1) showed increased affinity for INSL5 compared
with RXFP3 but no increase in GTP
S binding and little
change in
125
I R3/I5 binding. The reverse chimera, con-
sisting of RXFP4 (transmembrane domains, loops and
COOH-terminal tail) and the NH
2
-terminal region of
RXFP3 (CR13), had similar affinity to the wild-type
RXFP4 receptor for human relaxin-3, but lower affinity
for INSL5 (594), suggesting that the NH
2
terminus of
RXFP4 is important for INSL5 and possibly human re-
laxin-3 binding (TABLE 4).
The roles of the ECLs of RXFP3 were also investigated
using RXFP3 and RXFP4 chimeras. Chimeras of RXFP3,
with ECL1 or ECL3 replaced with the corresponding
loops of RXFP4, produced receptors that behaved essen-
tially like the RXFP3 wild-type receptor, suggesting that
ECL1 and ECL3 were not important in INSL5 binding to
either RXFP3 or RXFP4 (594). Receptors consisting of
RXFP3 with the ECL2 of RXFP4 (CR3) had higher af-
finity for INSL5, and a slightly lower affinity for human
relaxin-3, compared with RXFP3, similar to the findings
with CR1 (594). The potency of human relaxin-3 at CR3
was unchanged, whereas the potency of INSL5 was in-
creased (594). Receptors consisting of RXFP4 with the
ECL2 of RXFP3 (CR14) had similar affinity for human
relaxin-3, but lower affinity for INSL5 (594), suggesting
that ECL2 was important for INSL5 binding to RXFP4.
Functionally, CR14 behaved similarly to CR13, showing
a slight decrease in the potency of human relaxin-3 and a
larger decrease in the potency of INSL5 (594), suggesting
that the ECL2 of RXFP3 and RXFP4 are important for
ligand binding and may also have a role in activation
(TABLE 4).
Replacement of both the NH
2
terminus and ECL2 of
RXFP3 with the NH
2
-terminal region and ECL2 of
RXFP4 produced a chimeric receptor (CR5) that bound
human relaxin-3 and INSL5 in a similar manner to wild-
type RXFP4 (594), suggesting that both the ECL2 and
the NH
2
-terminal region of RXFP4 are important for
INSL5 binding. However, stimulation of chimeric CR5
with INSL5 does not increase GTP
S binding, suggesting
that the NH
2
terminus and ECL2 of RXFP4 (though
important for binding) are not sufficient for receptor ac-
tivation (TABLE 4). The reverse receptor, consisting of
RXFP4 with the NH
2
-terminal region and ECL2 replaced
with the equivalent regions of RXFP3 (CR15), had a
reduced affinity for INSL5 but no change in the affinity
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
423Physiol Rev VOL 93 JANUARY 2013 www.prv.org
for human relaxin-3. Correspondingly, while human re-
laxin-3 stimulation of CR15 increased GTP
S binding,
INSL5 was ineffective, suggesting that the NH
2
terminus
and ECL2 of RXFP4 are more important for ligand bind-
ing than for receptor activation (TABLE 4). Together,
these studies with chimeric receptors confirm that the
NH
2
-terminal region and ECL2 are both important for
human relaxin-3 and INSL5 binding, and thus may influ-
ence the activation of RXFP3 and RXFP4 receptors
(594).
The importance of the transmembrane (TM) spanning
regions of RXFP3 and RXFP4 was also investigated using
chimeras. Systematic replacement of the TM domains of
RXFP4 with those of RXFP3 showed that replacement
of TM3 or TM5 of RXFP4 with the corresponding regions
of RXFP3 (CR19 or CR21, respectively) caused a decrease
in affinity and a complete loss of INSL5 activity (594). This
suggests that TM3 and TM5 are involved in INSL5 binding
to RXFP4 and essential for activation. Although the reverse
chimera of RXFP3 with both TM3 and TM5 of RXFP4
(CR10) showed increased affinity for INSL5 (compared to
wild-type RXFP3), INSL5 still did not activate GTP
S bind-
ing (594). This suggests that TM3 and TM5 alone are not
enough to allow INSL5 activation of RXFP3. The chimera
of RXFP3 with TM2, TM3, TM5, and ECL2 of RXFP4
(CR12) exhibited similar affinity for both human relaxin-3
and INSL5 as the wild-type RXFP4 receptor (594), suggest-
ing that all of these regions were important for human re-
laxin-3 and INSL5 binding. Functionally, human relaxin-3
had the same potency for GTP
S binding at CR12 as the
wild-type RXFP4, and while INSL5 displayed increased po-
tency at CR12, it was still slightly lower than at RXFP4
(594). This suggests that TM2, TM3, TM5, and ECL2 are
all involved in ligand binding and activation of RXFP3 and
RXFP4 (TABLE 4).
Receptor Domain INSL5 INSL5
RXFP3 None 9.29 7.01 9.48 ND
CR1 8.84 7.99 9.10 ND
CR2 ECL1 9.17 7.03 9.21 ND
CR3 ECL2 8.94 7.94 9.02 ND
CR4 ECL3 9.19 6.92 9.32 ND
CR5 8.94 8.53 8.89 ND
CR8 8.67 8.58 8.88 8.54
CR12 8.79 8.29 9.08 8.45
RXFP4 None 8.81 8.66 9.01 8.94
CR13 8.95 7.66 9.11 7.74
CR14 ECL2 8.76 7.50 8.86 7.76
CR15 8.87 6.84 8.93 <7
CR16 9.04 7.06 9.09 <6
CR21 TM5 8.76 7.66 8.71 <6
Binding - pIC50 GTPγS binding - pEC50
Relaxin-3 Relaxin-3
N-term-TM5
N-term/ECL2
N-terminus
N-term-TM3
N-term-TM5
N-term/ECL2
N-terminus
TABLE 4. Functional domains of RXFP3 and RXFP4 receptors determined using receptor chimeras. RXFP3
residues are outlined in black, with RXFP4 in gray. Binding was determined using
125
I-labeled R3/I5 and
receptor activation by GTP
s binding. Data are from Reference 594.
BATHGATE ET AL.
424 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
IV. EXPRESSION AND SECRETION OF
RELAXIN FAMILY PEPTIDES IN
HUMANS AND OTHER SPECIES
A. Production and Secretion of Relaxin in
Mammals
Although relaxin was originally discovered on the basis of
its actions on the pregnant female reproductive tract in
lower species, it is clear that it also has additional roles in a
wide variety of tissues in the nonpregnant female and the
male. Thus, in those species where it has been examined,
relaxin is found and acts as a neuropeptide in brain and is
expressed in peripheral tissues such as heart, kidney, lung,
liver, and pancreas (FIGURE 4). In reproductive tissues, re-
laxin expression is found in the corpus luteum, uterus,
mammary gland, testis, and prostate. Whereas the general
distribution of relaxin in nonpregnant mammals points to
autocrine or paracrine functions, in pregnancy relaxin is
secreted in large amounts and has endocrine functions. The
source of this relaxin varies but can be the corpus luteum,
uterus, or placenta or combinations of these depending on
the species. The actions of relaxin during pregnancy are also
quite variable across species but are generally related to its
plasma secretion profile; thus in some species where levels
are high in late pregnancy it plays an important role in
modifying the reproductive tract in readiness for parturi-
tion, whereas in others where the plasma profile is different,
such as humans, its role is less clear.
The following sections outline the known sources of relaxin in
various species. We have not attempted to compare the rela-
tive expression of relaxin in different tissues due to the variety
of techniques that have been used to determine the sites of
expression. It should be noted that on occasion only one tech-
nique was used to demonstrate tissue expression, and in addi-
tion, the antibodies utilized were not always characterized for
cross-reactivity between different relaxin family peptides.
Therefore, some of these data may need to be confirmed using
independent techniques.
1. Detection of relaxin mRNA and protein in human
reproductive tissues
In nonpregnant women, human relaxin-2 mRNA expres-
sion in the ovary has been identified by Northern blotting,
RT-PCR, and Southern blotting (191, 246, 258). It is likely
that ovarian relaxin is produced in the corpus luteum since
human relaxin-2 is found in luteal extracts (391), luteal cyst
fluid (321), and in bathing fluid from dispersed corpora
lutea (451). Relaxin expression examined using immuno-
histochemistry reveals expression in the corpus luteum
(579), specifically to luteal cells (351), granulosa cells of
lutenized follicles (577), and cells from the lutenized theca
interna of developing follicles (61). In women, there is good
evidence that this relaxin is released into the circulation
during the menstrual cycle, since a small peak of relaxin
immunoreactivity occurs in the plasma 9–10 days follow-
ing ovulation (497).
Relaxin is also expressed in the secretory endometrium
(578, 579), and the glandular and luminal epithelia and
decidualized stromal cells of the endometrium (71). Relaxin
mRNA expression has been detected by RT-PCR in glan-
dular epithelial and stromal cells from cultures of human
endometrium (409). Both human relaxin-1 and human re-
laxin-2 mRNA were found by RT-PCR and qRT-PCR in
the uterus during the menstrual phase of the cycle (67).
In human pregnancy, the corpus luteum is the major source
of immunoreactive circulating relaxin (336, 391, 392). The
levels peak in the first trimester, and ovarian venous blood
levels are three to four times higher at 6 wk gestation than in
the luteal phase of the menstrual cycle (283). The circulat-
ing relaxin from the corpus luteum is human relaxin-2. The
corpus luteum contains human relaxin-2 mRNA (258), and
human relaxin-2 has been isolated from plasma of pregnant
women (324, 566, 567). Consistent with this being the ma-
jor source of relaxin in pregnancy, in women with ovarian
failure and those lacking corpora lutea who become preg-
nant by in vitro fertilization, circulating levels of relaxin are
undetectable (148, 269). Comparison of the pattern of re-
laxin secretion in human pregnancy with that in other spe-
cies reveals major differences. In pregnant women, relaxin
levels peak in early pregnancy at fairly modest levels but
then fall, whereas in other mammals such as pigs, rats, and
mice relaxin levels are much higher and rise throughout
pregnancy, falling just prior to parturition (47).
During pregnancy there is also expression of relaxin in the
placenta (191), with Northern blotting detecting relaxin
mRNA in the amnion, chorion, decidua parietalis, basal
plate, and placental trophoblast (439). Relaxin immunore-
activity was found in the amnion (70, 287, 439), chorion
(70, 287, 439), decidua (287, 579) (70, 439), basal plate
(166) (287, 439), and placental trophoblast (439, 579).
Relaxin has also been purified from the placenta (166, 570).
In other tissues, such as breast tissue, relaxin mRNA was
detected by RT-PCR and subsequent Southern analysis,
with increased expression in neoplastic samples (522). Re-
laxin immunoreactivity is present in lobular and ductal ep-
ithelium and myoepithelium of the normal and neoplastic
breast, with secretion detected from metaplastic epithelium
(355). The peptide is also present in milk (141).
In the male, similar approaches have identified human re-
laxin-1 and human relaxin-2 mRNA (191, 258) and relaxin
immunoreactivity in the prostate (489), specifically within
the glandular epithelium of the prostate, the glandular epi-
thelium of the seminal vesicles and the ampulla of the vas
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
425Physiol Rev VOL 93 JANUARY 2013 www.prv.org
deferens (576), and immunoreactive human relaxin-2 is
also secreted into seminal plasma (117, 320, 557, 567).
2. Detection of relaxin mRNA and protein in other
mammalian reproductive tissues
In rats and mice, as in humans, the corpus luteum is a major
source of relaxin particularly during pregnancy, but there
the resemblance ends. In ovaries from nonpregnant ani-
mals, relaxin mRNA (46, 80, 108, 190, 244) and protein
levels (473) are relatively low. Specifically, relaxin immu-
noreactivity is localized to the luteal cells of the ovary (14).
These levels increase at estrus (108) and in pseudopregnant
rats (473), but major increases only occur with pregnancy
(14, 472, 473). Unlike humans, rats and mice display a
marked increase in relaxin levels which peak just before
birth and then return to low levels post-partum (47).
In the nonpregnant mouse, relaxin mRNA expression has
been detected by Southern blotting of RT-PCR products in
the testis, epididymis, prostate, mammary gland, endome-
trium, and myometrium (46). Similarly in the rat, relaxin
mRNA expression has been detected by RT-PCR and
Northern blot analysis in the placenta, uterus, mammary
gland, testis, and prostate (190).
The source and function of relaxin in reproductive tissues
from other mammalian species have been described in detail
elsewhere (47). However, irrespective of the source, relaxin
levels are greatly increased during pregnancy in all mam-
mals that have been studied. In pigs, as in rats and mice, the
corpus luteum is the main source of relaxin in pregnancy,
but in the horse, rabbit, hamster, cat, and dog, the placenta
appears to be the main site of synthesis (see Ref. 47 for
details). The guinea pig is unusual in that the uterus is the
main site of relaxin production during pregnancy (47).
3. Relaxin as a neuropeptide in the CNS
There are no studies to date detailing expression of relaxin
within human brain. However, in the rat, relaxin mRNA
expression in the brain has been detected by RT-PCR (190)
and Northern blotting (402). Specific localization of the
peptide was also determined by in situ hybridization histo-
chemistry in anterior olfactory nuclei, taenia tecta, and piri-
form cortex (81, 333, 402), the orbital cortex (333, 402),
the fields CA1–2 of Ammon’s horn, the dentate gyrus of the
hippocampus and the neocortex (402), and the anterior
cingulated cortex and the arcuate nucleus (333) (FIGURE 6).
Immunohistochemistry has identified relaxin expression in
the cytoplasm and proximal processes of cell bodies in the
arcuate nucleus, anterior olfactory nucleus, taenia tecta,
and piriform cortex (333). In the mouse, relaxin expression
has been detected by Southern blotting of RT-PCR products
in the cortex, hippocampus, hypothalamus, thalamus,
pons/medulla, and cerebellum (46). Broadly the distribu-
tion of expression of relaxin in mouse brain resembles that
of the rat (306).
4. Distribution of relaxin in other tissues
In the human, relaxin has been detected in the circulation
using ELISA (497) and radioimmunoassay (141, 269, 391,
393, 426, 498) techniques. Cultures of human luteinizing
granulosa cells also secrete relaxin, as determined using an
immunoassay (499). In mice, relaxin expression has been
detected by Southern blotting of RT-PCR products in the
thymus, heart, and kidney, with lower levels in the lung,
spleen, and skin (46). RT-PCR identified relaxin expression
specifically in the medulla and cortex of the kidney (445),
and in the atria and ventricles of the heart (136). Radioim-
munoassay identified relaxin in the peripheral circulation in
pregnant mice (393). Distribution of relaxin expression in
the rat is similar to mice, with mRNA demonstrated by
RT-PCR in the kidney, heart, liver, and pancreas (190).
Radioimmunoassay has identified relaxin in the peripheral
circulation of pregnant rats (393, 472), and a reverse hemo-
lytic plaque assay detected relaxin that was secreted from
cultured atrial cardiomyocytes from neonatal rats (524). In
some instances, only relaxin-3 but not relaxin expression
was identified in the heart by RT-PCR (290, 443).
B. Production and Secretion of INSL3 in
Mammals
INSL3 expression occurs mainly in the reproductive system
in mammals with the highest expression in the testis. How-
ever, INSL3 is expressed at equivalent levels in the ruminant
ovary and has an expression pattern during pregnancy that
is similar to relaxin (40, 41, 432). As ruminants lack a
relaxin gene (563), it has been postulated that INSL3 may
act in lieu of relaxin in bovine and ovine species (48). As the
expression of INSL3 in all mammalian species has been
covered in detail elsewhere (47), this section will focus in
INSL3 in humans and rodents.
1. Detection of INSL3 mRNA and protein in the
mammalian reproductive system
INSL3 mRNA is highly expressed in the testis of every
mammalian species so far examined. In humans, INSL3
mRNA has been identified in the testis by Northern blot
analysis (2, 84), RT-PCR (254), and Western blotting
(227). Expression of INSL3 has been specifically localized
to the Leydig cells using in situ hybridization (254, 285) and
immunohistochemistry (24, 285) (FIGURE 4, TABLE 1).
INSL3 has also been detected by immunoassay in human
serum (8, 11, 50, 72, 172, 173), with higher levels in males
compared with females correlating with the high expression
in the Leydig cells of the testis. In the male, INSL3 levels in
the peripheral circulation decline with age (11). Immunoas-
say has also identified INSL3 expression in the amniotic
BATHGATE ET AL.
426 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
fluid of male, but not female, fetuses (10) suggesting that
INSL3 is expressed in fetal testis as has been demonstrated
in other mammals (47) which also correlates with its role in
testis descent (see sect. VIIB). INSL3 mRNA and protein are
also expressed in the prostate, with strong expression in the
basal epithelium (286). In the female, INSL3 is expressed in
the ovary, specifically within the corpus luteum as deter-
mined by RT-PCR (521), Northern blotting (521), and im-
munohistochemistry (24), in addition to expression in the
trophoblast of the ovary determined using RT-PCR (521)
and the theca interna cells using immunohistochemistry
(24). Both in situ hybridization and immunostaining have
identified INSL3 expression in the endometrium, specifi-
cally within the epithelial cells, endometrial gland cells, and
stromal cells (230). The peptide is expressed in human
breast tissue, determined using RT-PCR, with in situ hy-
Cx
Cells positive for relaxin
mRNA and/or immunoreactivity
Cells positive for relaxin-3
mRNA and/or immunoreactivity
Main paths of relaxin-3
immunoreactive fibers
Subsidiary paths of relaxin-3
immunoreactive fibers
CC
CPut
Orb
OB
AON
AON
VTT
DTT
Acb
GP
VP
DBB
Amy AHi
SON
MPA
PVN DMH
PHLH
VMH
Arc
ME
AP
IL
NL
IPN
SuM
SN VR
PnR
NI
DTg
PBN
DR
PAG
ND
PP
IGL
Re
Th
CM
APT
SFO
PVA
DG
Hi
Sub
SC IC
Pir
Hpt
SHy
BST
S
LS
C
NST ARP
Sp5
IO
Amb
Relaxin
binding sites
RXFP3
binding sites
Regions containing both relaxin
and RXFP3 binding sites
FIGURE 6. Localization of relaxin-RXFP1 and relaxin-3-RXFP3 sites in rat brain. The figure shows a sagittal
section of rat brain with indications of the areas containing cells staining positive for relaxin mRNA or
immunoreactivity or relaxin-3 mRNA or immunoreactivity. The regions containing RXFP1 binding sites deter-
mined using radiolabeled relaxin (blue) or RXFP3 binding sites obtained using labeled
125
I-R3/I5 (yellow) are
outlined together with sites containing both receptors (green). Acb, nucleus accumbens; Ahi, amygdalohippo-
campal area; Amb, nucleus ambiguus; Amy, amygdala; AON, anterior olfactory nucleus; AP, anterior pituitary;
APT, anterior pretectal nucleus; Arc, arcuate nucleus of hypothalamus; ARP, area postrema; BST, bed
nucleus of stria terminalis; C, cerebellum; CC, corpus callosum; CM, central medial thalamic nucleus; CPut,
caudate putamen; Cx, cerebral cortex; DBB, diagonal band of Broca; DG, dentate gyrus; DMH, dorsomedial
nucleus of hypothalamus; DR, dorsal raphe nucleus; DTg, dorsal tegmental nucleus; DTT, dorsal taenia tecta;
GP, globus pallidus; Hi, hippocampus; Hpt, hypothalamus; IC, inferior colliculus; IGL, intergeniculate leaflet; IL,
intermediate lobe of pituitary; IO, inferior olive; IPN, interpeduncular nucleus; LH, lateral hypothalamus; LS,
lateral septum; ME, median eminence; MPA, medial preoptic area; ND, nucleus of Darkschewitsch; NI,
nucleus incertus; NL, neural lobe of pituitary; NST, nucleus of solitary tract; OB, olfactory bulb; Orb, orbital
cortex; PAG, periaqueductal gray; PBN, parabrachial nucleus; PF, posterior hypothalamus; Pir, piriform
cortex; PnR, pontine raphe; PP, peripeduncular nucleus; PVA, paraventricular thalamic area; PVN, paraven-
tricular nucleus of hypothalamus; Re, reuniens thalamic nucleus; S, septum; SC, super colliculus; SFO,
subfornical organ; SHy, septohypothalamic nucleus; SN, substantia nigra; SON, supraoptic nucleus; Sp5,
spinal trigeminal tract; Sub, subiculum; SuM, supramammillary nucleus; Th, thalamus; VMH, ventromedial
nucleus of hypothalamus; VP, ventral pallidum; VR, ventral raphe nuclei; VTT, ventral taenia tecta. Sources of
material: relaxin mRNA and IR (81, 402); piriform cortex; septohypothalamic nucleus; arcuate nucleus;
horizontal/vert diagonal band of Broca; median preoptic region; 3rd ventricle lining (not included in schematic);
paraventricular hypothalamic nucleus; paraventricular thalamic area; orbital cortex and neocortex; anterior
olfactory area; taenia tecta (dorsal and ventral); anterior olfactory nucleus; taenia tecta; piriform; hippocam-
pus; CA1–3 of Ammons horn and dentate gyrus; orbital cortex/neocortex; relaxin binding (81, 333, 402);
relaxin-3-positive IR neurons and fibers (329, 508); RXFP3 binding sites and mRNA (486, 487, 507); inferior
olive; taenia tecta; ventral pallidum; diagonal band of Broca; septohypothalamic; lateral, posterior, dorsomedial
hypothalamus; intergeniculate nucleus; peripeduncular nucleus.
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
427Physiol Rev VOL 93 JANUARY 2013 www.prv.org
bridization showing localization of INSL3 to the tubu-
loaveolar and ductal epithelium (227). INSL3 expression
has also been detected in placental tissue using Western
blotting (227, 230), with specific expression in the tropho-
blasts of anchoring villi and fetal membranes, maternal
blood vessels, and the basal plate identified using in situ
hybridization and immunostaining (230).
In the mouse and the rat, INSL3 is expressed in the testis as
determined by Northern blotting (23, 412, 491, 595) and
RT-PCR (23, 276), with specific expression within the Ley-
dig cells recorded using Northern blotting (424), in situ
hybridization (23, 412, 424, 525), and immunohistochem-
istry (23, 184, 412). Levels of INSL3 mRNA are highest in
the fetal testis shortly before birth, after which they decrease
until puberty when they again rise.
In the rat, INSL3 has been identified in the circulation of the
male using immunoassay but levels in the female are very
low (9, 68). Plasma levels in rats are correlated with mRNA
levels being highest just before birth and during adulthood
(9, 68) and decline with age as in humans (9). There is also
evidence for the transport of INSL3 across the blood-testis
barrier into the germinal compartment (9). In the mouse,
INSL3 levels in the blood are also much higher in males
than females (9). INSL3 protein expression has also been
demonstrated in the seminiferous tubules of the (9) mouse
using immunohistochemistry (23). RT-PCR has detected
expression of INSL3 mRNA in the epididymis (23), pros-
tate (23), and the posterior part of male embryos (595). In
the female mouse and rat, INSL3 mRNA was detected in
the ovary using Northern blotting (491, 595) and RT-PCR
(23, 276), with specific expression in the mouse in the cor-
pus luteum and stromal cells determined using immunohis-
tochemistry (23). In addition, in the rat, the use of RT-PCR
shows specific expression in the thecal cells surrounding
preovulatory follicles (276).
2. Is INSL3 a neuropeptide?
Despite evidence for the expression of the INSL3 recep-
tor, RXFP2, in brain, no studies have identified INSL3
expression in the CNS of rodents or humans (189, 469).
Circulating levels of INSL3 could potentially affect some
areas of the brain such as the pituitary, hypothalamus,
and areas of the brain stem, although this remains to be
demonstrated (86).
3. Detection of INSL3 in other peripheral tissues
There is limited evidence for the expression of INSL3 in
nonreproductive tissues. In humans, INSL3 expression, as
determined using in situ hybridization, is upregulated in the
neoplastic thyroid gland (228).
C. The Location and Release of Relaxin-3 in
Mammals
1. Identification of relaxin-3 in brain
In contrast to the widespread expression of relaxin and
INSL3, studies that have mapped the specific distribution of
relaxin-3 and RXFP3 expression have provided valuable
insights into the physiological functions of this ligand-re-
ceptor pair. Relaxin-3 expression measured by Southern
blotting of RT-PCR products from mouse tissue (46)
showed relaxin-3 mRNA in abundance in the brain (FIGURE
6); at moderate levels in the thymus, kidney, and spleen; and
at low levels in the heart and liver, with no noticeable gen-
der differences. Northern blots and in situ hybridization
studies showed that relaxin-3 is localized to cells of the pars
ventromedialis of the dorsal tegmental nucleus (46). The
expression of rat relaxin-3 was subsequently found to mir-
ror that of the mouse homolog (FIGURE 6) (80). Similarly,
relaxin-3 expression is highest in the human brain, although
substantial expression is also found in the testis, where its
role remains to be demonstrated (314). The high levels of
relaxin-3 expression in the brain of numerous species have
led to a focus on the role of the peptide in the CNS.
Most neuroanatomical studies have been conducted in ro-
dents, and in rat brain, the distribution of relaxin-3 has
been characterized with two different antibodies (329,
516). It is expressed in the nucleus incertus (NI), located in
the median dorsal pons and also referred to as the dorsal
tegmental nucleus pars ventromedialis (373), nucleus reces-
sus pontismedialis (261), or nucleus “O” (413). Neurons
expressing relaxin-3 are also present in the pontine raphe
nucleus; the anterior, lateral, and ventrolateral periaque-
ductal gray (PAG); and in an area dorsal to the lateral
substantia nigra (FIGURE 6) (329). As in the rat, the primary
source of relaxin-3 in the mouse brain is the NI, with some
relaxin-3-positive neurons also located in the pontine cen-
tral gray and adjacent to the fourth ventricle (487). Discrete
populations of relaxin-3-containing neurons were also ob-
served in the pontine raphe nucleus, anterior periaqueduc-
tal gray, and a region dorsal to the substantia nigra and
medial to the peripeduncular nucleus. A recent study of sites
of relaxin-3 expression in zebrafish has revealed expression
patterns similar to those observed in rodents (125). Relax-
in-3 expression occurs in neurons located near the fourth
ventricle and in the periaqueductal gray (confirmed by co-
localization with proenkephalin-like gene 1). In the ma-
caque (Macaca fascicularis) brain, relaxin-3 is expressed in
areas similar to the NI in rodents (332). In the rhesus ma-
caque (Macaca mulatta), using a different antibody, there
was a similar pattern of expression (478). This consistent,
specific and restricted expression of relaxin-3 in multiple
species suggests that relaxin-3 has quite conserved and thus
important functions in the brain.
BATHGATE ET AL.
428 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
The target nuclei of the neuronal projections from relaxin-
3-expressing cell bodies provide further insights into relax-
in-3 function. The projections of NI neurons in the rat brain
have been extensively mapped (187, 395). Ascending pro-
jections from the NI connect to the raphe, mammillary com-
plex, hippocampus, cortex, amygdala, habenula, medial
septum, and hypothalamus (FIGURE 6). Overall, the NI ap-
pears to have strong reciprocal connections with the median
raphe and the interpeduncular nuclei. The NI appears to
have important roles in behavioral arousal (187, 395).
D. Distribution of Other Relaxin Family
Peptides in Mammals
INSL4 is highly expressed in the placenta, and RT-PCR
analysis suggests that it is most highly expressed in the
human placenta during the first trimester of pregnancy (94).
No INSL4 transcripts were identified in human ovary, pros-
tate, testis, brain, skin, esophagus, adrenal, breast, kidney,
thyroid, stomach, liver, colon, rectum, omentum, or spleen.
Immunoreactive INSL4 is detected in the plasma and amni-
otic fluid during pregnancy, with higher amniotic fluid lev-
els found in abnormal pregnancies (such as trisomy 21),
suggesting that INSL4 may be a marker of chromosomal
abnormalities during pregnancy (368, 369).
The INSL5 gene was identified from colon EST libraries,
and the highest expression of INSL5 mRNA is in the colon
in mice and human (100). INSL5 mRNA is also expressed
in the kidney, thymus, heart ovary, and brain (FIGURE 4),
although there are some differences in tissue expression
between the mice and humans, with mRNA also detected in
human uterus, placenta, prostate, spleen and bone marrow,
and in mouse testis (100, 236, 315). Relatively little is
known about the distribution of INSL5 protein. In the kid-
ney, INSL5 appears to be located in the cytoplasm of a
subset of epithelial cells in the loop of Henle (236). In the
mouse brain, putative INSL5 immunoreactive cells are lo-
cated in the paraventricular, supraoptic, accessory secre-
tory and supraoptic retrochiasmatic nucleus throughout the
rostrocaudal extent of the hypothalamus and pituitary
(137). In contrast, another study showed a lack of INSL5
mRNA expression in the mouse brain (507).
INSL6 mRNA was identified by Northern blotting only in
the testis of human tissues and not in heart, brain, placenta,
lung, liver, skeletal muscle, kidney, pancreas, spinal cord,
lymph node, trachea, adrenal gland, or bone marrow (317).
The distribution of rat INSL6 mRNA was similar as deter-
mined by Northern blotting, with mRNA identified in the
testis and prostate of rat tissues (317). In situ hybridization
of rat and Rhesus macaque tissues identified INSL6 mRNA
in the spermatocytes and round spermatids in the seminif-
erous tubules but not in the Leydig cells (317). Studies in
mice have also demonstrated that INSL6 is expressed in
meiotic and postmeiotic germ cells correlating with the
marked reduction in sperm numbers and motility in INSL6
knockout mice (85). A recent study has demonstrated that
INSL6 is expressed in skeletal muscle in mice and may func-
tion as a myogenic regenerative factor (584).
V. LOCALIZATION OF RELAXIN FAMILY
PEPTIDE RECEPTOR MRNA AND
PROTEIN IN MAMMALIAN TISSUES
The tissue distribution of receptors for relaxin family pep-
tides has been determined by their mRNA expression pat-
terns using RT-PCR, Northern blotting, or in situ hybrid-
ization or by protein expression using immunohistochem-
istry or receptor autoradiography (FIGURE 6;TABLE 1). The
expression patterns of all the receptors generally match
their identified roles as cognate receptors for their respective
ligands. There is much information regarding RXFP1 ex-
pression, both in terms of the variety of approaches used
and the tissues examined (TABLE 1), with receptor expres-
sion being identified in both female and male reproductive
tissues, the brain and numerous nonreproductive tissues
such as the kidney, heart, and lung. Less information is
available for RXFP2, RXFP3, and RXFP4, although more
detailed evidence is now emerging for RXFP3 expression in
the brain, suggesting that this receptor may have promise as
a therapeutic target (see sect. VIIIC).
A. The Tissue Distribution of RXFP1:
the Receptor for Relaxin
1. Reproductive tissues
Prior to the discovery of RXFP1, localization of the un-
known relaxin receptor was determined by monitoring the
binding of labeled relaxin peptides to various tissues. In this
manner, early studies demonstrated relaxin binding in ho-
mogenates of mouse pubic symphysis and uterine tissue, rat
mammary gland and guinea pig pubic symphysis and cervix
(TABLE 1) (363). Other autoradiographic studies utilised a
relaxin variant (B33 relaxin, which facilitates peptide label-
ing with
32
Por
33
P) to identify relaxin binding in rat uterus
(401, 403, 515). Following receptor deorphanization,
RXFP1 distribution was identified using immunohisto-
chemistry in both myometrial and epithelial cells of the rat
uterus (239). In humans, RXFP1 has been identified by
binding and immunohistochemistry in luminal and glandu-
lar epithelial cells, myometrium, blood vessels, and stromal
extracellular matrix (TABLE 1) (288). Similar localization
has been reported using
33
P relaxin (67) or immunohisto-
chemistry (297, 325). Other immunohistochemical studies
for RXFP1 have reported binding solely to endometrial
stromal cells in humans but also to glandular epithelial cells
of the marmoset endometrium (255). There is some evi-
dence to suggest that RXFP1 expression in the human en-
dometrium varies with the menstrual cycle. Autoradiogra-
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
429Physiol Rev VOL 93 JANUARY 2013 www.prv.org
phy was used to localize specific relaxin binding to human
uterus and showed binding to the glandular epithelium and
endometrial luminal epithelium. Both binding and RXFP1
mRNA levels increased markedly in the early secretory
phase of the menstrual cycle compared with the prolifera-
tive phases (67). Weaker binding occurred in endometrial
stromal tissue and little in myometrium (67). Immunohis-
tochemical studies are more equivocal with either broad
agreement (297) with binding studies or evidence for more
intense staining during the proliferative phase of the cycle in
studies with a high degree of variability (325).
In human breast tissue, RXFP1 was detected in the stromal
mesenchyme adjacent to the glandular epithelia and in epi-
thelial-derived cells in some breast tumors (255). Cultured
slices of ovarian cortical tissue displayed RXFP1 mRNA
and protein in flat follicular cells of primordial follicles and
granulosa cells of primary and secondary follicles (474).
The distribution of RXFP1 in rodent reproductive tissues
is similar to that in humans. RXFP1 mRNA has been
detected by RT-PCR in the mouse uterus and testis (274).
Incorporation of the LacZ reporter gene into the RXFP1
knockout mouse allowed receptor expression to be dem-
onstrated near the spermatids and in the Leydig cells of
the testis, the prostate, the nipple, the circular layer of the
myometrium, the lamina propria under the luminal epi-
thelium and the longitudinal layer of the myometrium,
the basal layer of the vaginal epithelium and vaginal
smooth muscle cells, as well as vascular supporting tissue
and vascular epithelium of the oviduct (293). In rats,
antibodies directed against the ectodomain of RXFP1
identified receptor protein in the myometrial and the ep-
ithelial layer of the endometrium of the uterus, the mus-
cularis layer of the vagina, and the myometrial layer of
the cervix (239). RXFP1 has also been detected in rat
nipple and mammary gland (300).
2. RXFP1 receptors in the CNS
RXFP1 receptors identified by receptor autoradiography
are widely distributed in the brain localized to discrete re-
gions of the olfactory system, neocortex, hypothalamus,
hippocampus, thalamus, amygdala, midbrain, and medulla
(400, 404, 515) (FIGURE 6). High levels of RXFP1 binding
in the subfornical organ (SFO), organum vasculosum of the
lamina terminalis (OVLT), and the paraventricular and su-
praoptic hypothalamic nuclei (400, 404, 515) provide the
anatomical and biochemical basis for the control of plasma
osmolality by relaxin (361, 362, 483, 504, 505). RXFP1
mRNA has also been demonstrated in rat brain by North-
ern blotting (237, 455) and by RT-PCR (167, 290, 298,
443), in mouse brain using a reporter gene approach (293)
and by Northern blotting (455), and in human brain using
RT-PCR (239).
3. Localization of RXFP1 in the cardiovascular and
renal systems
The important roles of relaxin in the cardiovascular adap-
tive changes associated with pregnancy are reflected in the
presence of RXFP1 mRNA in the rat (237) and human
kidney (239). In rodents, the heart is also clearly influenced
by relaxin. Atria of both male and female rats contain a high
density of relaxin binding sites (401, 402, 515), and RXFP1
mRNA has been detected in the rat (237, 290) and mouse
(293) heart. However, lower levels of RXFP1 mRNA have
also been detected in the human heart (TABLE 1) (239).
More recent studies using Western blotting show RXFP1
expression in left atrium but not left ventricle in both failing
and nonfailing human heart (127). Relaxin is a potent chro-
notropic and inotropic agent in the rat heart (271, 290, 417,
418) and also has inotropic effects on the human heart
(127) (see sect. VIIA). It produces its effects by acting di-
rectly on RXFP1 receptors located mainly in the atria.
B. Tissue Distribution of RXFP2:
the Receptor for INSL3
1. Localization of RXFP2 in reproductive tissues
Although fewer studies have examined the localization of
RXFP2 compared with RXFP1, it is clear that the RXFP2/
INSL3 receptor/ligand pairing has highly specialized roles
in reproduction. Studies of the distribution of RXFP2 show
mRNA in rat ovary and in rat and mouse testis and guber-
naculum by Northern analysis, RT-PCR, and in situ hybrid-
ization (12, 276, 298, 301, 453) (FIGURE 4;TABLE 1). In the
rat, RXFP2 receptor mRNA has been localized to the
oocyte in the ovary and the germ cells of the seminiferous
tubules associated with germ cell function (276) (See sect.
VIIB). A more recent study has confirmed the expression of
RXFP2 mRNA in rat postmeiotic germ cells and also local-
ized RXFP2 immunoreactivity to human postmeiotic germ
cells (12). This study also demonstrated RXFP2 mRNA in
rat Leydig cell, immunoreactive RXFP2 in human Leydig
cells, and INSL3 binding in a Leydig cell line. In human,
RXFP2 mRNA is found in uterus and testis (239, 354). The
pattern of localization of RXFP2 thus correlates with the
known roles of INSL3/RXFP2 in gubernacular develop-
ment and reproductive physiology (see TABLE 1 and sect.
VIIB).
2. Is there a role for RXFP2 in the CNS?
RXFP2 mRNA is expressed in the rat brain with high den-
sities in the thalamus where it is present in the intralaminar,
parafascicular, dorsolateral, ventrolateral, and posterior
thalamic nuclei (460). Mapping of the RXFP2 protein using
125
I INSL3 revealed high peptide binding in thalamic re-
gions but also in the striatum and certain layers of the
cortex, suggesting axonal transport of receptors to terminal
BATHGATE ET AL.
430 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
sites (460). However, in situ hybridization studies of the
distribution of INSL3 mRNA using oligo probes failed to
provide evidence for INSL3 expression in rat brain (469),
perhaps suggesting that central RXFP2 receptors respond
to circulating INSL3 (see sect. IVB). Comparison between
RXFP2 expression patterns in the rat and mouse revealed
detailed differences, most notably the presence of both
RXFP2 mRNA and protein in the cortex of the mouse with
mRNA found in layer V and layer VIb and
125
I INSL3
binding present in deeper layers and also layer I (189).
However, as in the rat, there was no evidence for INSL3
mRNA expression in mouse brain. (189).
3. Evidence for RXFP2 in other peripheral tissues
In the initial study that described the deorphanization of
RXFP1 (LGR7) (239), examination of the tissue distribu-
tion of the related RXFP2 receptor by RT-PCR provided
evidence for mRNA expression in brain, kidney, muscle,
testis, thyroid, uterus, bone marrow, and peripheral lym-
phocytes (FIGURE 4;TABLE 1). Signal strength was strong in
testis; weaker in lymphocytes, bone marrow, and brain; and
weak in the remaining tissues (239). In a more detailed
study in rat kidney, both RXFP2 mRNA and
125
I INSL3
binding were found in renal glomeruli and mesangial cells,
with particularly high levels observed in late stage gestation.
RXFP2 receptor-mediated responses are present in human
and mouse osteoblasts related to roles for INSL3 in bone
function (162) (see sect. VIIB). RXFP2 mRNA is also ex-
pressed in human thyroid tissues and in mouse follicular
thyroid epithelial cells (225).
C. Localization of RXFP3 and RXFP4: The
Receptors for Relaxin-3 and INSL5
1. RXFP3 has a widespread distribution in the CNS
Following deorphanization, RXFP3 was found to be highly
expressed in the human brain as determined by RT-PCR
(352). Subsequent studies identified the brain as the pre-
dominant site of RXFP3 expression in the mouse by North-
ern blotting and in both mouse and rat by in situ hybridiza-
tion (62, 314). In rat brain, high levels of mRNA were
found in the olfactory bulb, paraventricular and supraoptic
nuclei, and preoptic and posterior areas of the hypothala-
mus, hippocampus, septum, and amygdala with lower lev-
els in cortex, peraqueductal grey, nucleus incertus, and ar-
eas of brain stem (FIGURE 6) (314, 508). The chimeric pep-
tide
125
I-R3/I5, a selective, high-affinity ligand for RXFP3
and RXFP4, has been used to identify chimeric receptor
binding sites in the cerebral cortex, olfactory bulb, and
superior colliculus in the rat (RXFP4 is a pseudogene in the
rat) (FIGURE 6) (312).
Recent studies provide a comprehensive description of the
distribution of RXFP3 in mouse brain (487) by both in situ
hybridization and peptide binding. While the overall pat-
tern is very similar to that in the rat, suggesting a strongly
conserved ligand-receptor pairing, there are some detailed
differences. For example, in the mouse, both peptide and
RXFP3 are present in the substantia innominata and the
olivary and posterior pretectal nuclei, whereas in the rat
expression in the olfactory bulb, entorhinal cortex, arcuate
nucleus, and paraventricular thalamic nucleus is more dom-
inant (487). As yet, there are no systematic studies of
RXFP3 localization in human or primate brain.
2. Evidence for RXFP3 in peripheral tissues
The first studies on the expression of RXFP3 mRNA in
human tissues indicated very low expression of RXFP3
mRNA in the adrenal gland, testis, salivary gland, and pan-
creas by RT-PCR. Subsequent studies demonstrated
RXFP3 mRNA expression in the human testis by RT-PCR
(314), although the function of the receptor in this tissue is
not known.
3. Expression of RXFP4
Studies prior to the deorphanization of RXFP4 examined
the distribution of the orphan GPCR, GPR100 using
Northern blotting of human tissues. Expression of GPR100
was identified principally in peripheral tissues including
heart, skeletal muscle, salivary gland, bladder, kidney, liver,
placenta, stomach, jejunum, thyroid, ovary, and bone mar-
row with the highest expression in the pancreas (63) (FIG-
URE 4;TABLE 1). Subsequent studies following the deorpha-
nization of RXFP3 showed high levels of receptor mRNA
expression by RT-PCR in the human colon with additional
expression in the placenta, testis, thymus, prostate, kidney,
and brain (313). The expression of RXFP4 in the colon
matches the expression of the INSL5 peptide and indicates
potential functions of INSL5 as a gut hormone (see sect.
VIID).
VI. SIGNAL TRANSDUCTION PATHWAYS
ACTIVATED BY RELAXIN FAMILY
PEPTIDE RECEPTORS
A. Pleiotropic Signaling Pathways Activated
by RXFP1
Since type A LGRs and the coevolved genes encoding gly-
coprotein hormone subunits can be traced to both nema-
todes and insects, the three subfamilies of LGRs are thought
to have evolved before the emergence of vertebrates and nem-
atodes (237). The signaling pathways activated by RXFP1 and
RXFP2 therefore represent one of the earliest forms of GPCR
signaling (42). RXFP1 and RXFP2 couple to a variety of G
proteins to influence cAMP accumulation within a number of
cell lines (198, 237, 239). In addition, there is evidence for
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
431Physiol Rev VOL 93 JANUARY 2013 www.prv.org
activation of MAP kinases, tyrosine kinases and nitric oxide
(NO) as well as activation of signaling pathways associated
with connective tissue metabolism, much of which predates
receptor identification (FIGURE 7A).
There is further evidence suggesting that relaxin activates
the glucocorticoid receptor (GR), a nuclear receptor that
acts as a ligand-dependent transcription factor (130). Acti-
vation of the GR by relaxin, and the subsequent changes in
gene transcription, may account for the many effects of
relaxin on the expression levels of a variety of proteins,
including those involved in connective tissue metabolism
(see sect. VIA4). Relaxin stimulation of differentiated
THP-1 cells (into a macrophage phenotype) blunts the pro-
duction of cytokines including interleukin (IL)-1, IL-6, and
tumour necrosis factor (TNF)-
, and this effect is abolished
by the GR antagonist RU486 (128). Relaxin coimmunopre-
cipitates with the GR, and the amount of the GR within the
nucleus increased after 30 min stimulation with relaxin
(128). Furthermore, a modified relaxin that was unable to
activate RXFP1 could still interact with the GR, and relaxin
was found to cause phosphorylation of S211 of the GR,
which is used as a biomarker of agonist-related receptor
activation (128). Indeed relaxin, via its interaction with the
GR, is also able to autoregulate its own expression (131).
Thus, although relaxin appears to interact with the GR,
independently of its own receptor, the relative contribution
of this pathway to the physiological effects of relaxin, and
how this integrates with the more classical relaxin-stimu-
lated signaling pathways, remains to be determined.
1. Increases in cAMP levels mediated by RXFP1:
a second messenger system influenced by many
interacting mechanisms
In studies predating receptor identification, treatment with
relaxin was found to cause an increase in cAMP accumula-
tion in THP-1 cells (411), MCF-7 cells (57), the mouse
pubic symphysis (69), uterine strips (446), uterine longitu-
dinal muscle (399) from estrogen-primed rats, and in cul-
tures of human endometrial cells (156), human endometrial
glandular epithelial cells (96), newborn rhesus monkey
RX F P 2 AC
ATP
cAMP
βγ
B
CRE Transcription
INSL3
RXFP2 Signalling
AC
RXFP4
D
RXFP4 Signalling
PLC
DAG
INSL5
AC
RXFP3
C
RXFP3 Signalling
PLC
PI3K
AP1 Transcription
Src
Shc
SOS
Transactivation
Grb
EGFR
Ras
Raf
MEK1/2
ERK1/2
NFκB Transcription
PKC IκB
Relaxin-3
RXFP1
AC5
Gαi3
AC
ATP
cAMP
ATP
cAMP
βγ
PI3K
PKCζ
A
NFκB Transcription CRE Transcription
Relaxin
RXFP1 Signalling
βγ
Akt
ERK1/2
NO
PIP2
InsP3
Ca2+
Gαs
GαoB
Gαs
GαoB
GαoB
Gαi2 Gα16 GαoB
Gαi2
FIGURE 7. Signal transduction mechanisms activated by interaction of relaxin family peptides with their
cognate RXFP receptors. Relaxin interacts with RXFP1 (A) to cause coupling of the receptor with G
sto
activate adenylyl cyclase (AC) and G
OB
to modulate this effect (198, 236, 237). The receptor also couples to
G
i3
and
␤␥
subunits from this interaction to activate PI-3-kinase to cause translocation of PKC-
, which in turn
activates AC5 (198, 204, 382, 383). cAMP resulting from these interactions appears to be compartmen-
talized and influences different signaling outcomes (197). INSL3 activates RXFP2 (B) to couple to G
s
and G
OB
and influence cAMP production (198, 301). Relaxin-3 binds to RXFP3 (C) causing coupling to G
i2
and G
OB
and inhibition of AC (314). Release of
␤␥
subunits from these G proteins activates MAPK signaling and NF
B
transcription (545, 546). Relatively little is known of INSL5/RXFP4 signaling (D), but the receptor couples to
G
i2
and G
OB
in a similar fashion to RXFP3 (313) but also couples to G
16
and Ca
2
when this promiscuous
G protein has been transfected into cells (315).
BATHGATE ET AL.
432 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
uterine cells (294), rat myometrial cells (235), and rat ante-
rior pituitary cells (109). The physiological relevance of the
cAMP response to relaxin is particularly well demonstrated
in human endometrial stromal cells, where basal and relaxin-
stimulated cAMP levels are enhanced by inhibition of phos-
phodiesterase (PDE) 4 (FIGURE 7A), and relaxin stimulation
and PDE4 inhibition act synergistically to induce decidual-
ization (37). Increases in cAMP in response to relaxin (243,
512) or addition of a cell-permeable cAMP analog (517)
induce decidualization, which is an important process re-
quired to support implantation of the developing embryo.
Increases in cAMP are also linked to the physiological ef-
fects of relaxin upon angiogenesis; treatment with human
relaxin in a murine model increased the degree of angiogen-
esis at wound sites, an effect that was associated with an
increased expression of vascular endothelial growth factor
(VEGF), an important proangiogenic protein (see sect.
VIIA4) (541). Interestingly, in cultures of normal human
endometrial cells (NHE cells), human relaxin increased
VEGF expression, and these effects are prevented by AC
inhibition and mimicked by either the AC activator for-
skolin or a PDE inhibitor (540). This suggests that relax-
in-stimulated cAMP production also mediates increased
VEGF transcription, and thus angiogenesis (see sect.
VIIA4).
Following receptor characterization, the importance of
cAMP as a signaling pathway for relaxin was cemented by
the observation that constitutively active mutants of both
RXFP1 and RXFP2 (transmembrane helix 6: D637Y) in-
creased cAMP accumulation in a ligand-independent man-
ner (FIGURE 3) (237, 239). Many subsequent studies have
focused on cAMP accumulation generated by activation of
these two receptors. It is now well recognized that both
RXFP1 and RXFP2 couple to G
s
to increase cAMP (198,
237, 239), an effect that is negatively modulated by cou-
pling to G
OB
(198). Only RXFP1 can couple to G
i3
to
activate a further surge of cAMP accumulation via a G
␤␥
-
phosphatidylinositol 3-kinase (PI3K)-protein kinase C
(PKC)-
pathway to stimulate adenylyl cyclase 5 (AC5) (FIG-
URE 7A)(198, 204, 382, 383).
Activation of the unique G
i3
pathway by RXFP1 was initially
identified in THP-1 cells that endogenously express the recep-
tor. In this cell line, relaxin causes a biphasic increase in cAMP
accumulation over time, with the later phase partially suscep-
tible to the PI3K inhibitors LY294002 and wortmannin;
subsequently, relaxin stimulation of RXFP1 was also
shown to increase PI3K activity (383). PKC-
was proposed
as the candidate protein that linked PI3K activation to
cAMP formation. Indeed, studies that predated receptor
identification in human cultured secretory endometrial
stromal cells showed that human relaxin treatment over 4
days increased the amount of PKC activity in membranes
versus cytosolic fractions of the cell (272); these observa-
tions support the concept of relaxin-stimulated transloca-
tion of PKC to the cell membrane.
The atypical PKC isoforms, including PKC-
, are insensitive
to diacylglycerol and Ca
2
but are either directly or indi-
rectly activated by phosphatidylinositol 3,4,5-trisphos-
phate (PIP
3
) and other lipids (378, 397, 493). Furthermore,
some general PKC inhibitors (including bisindoylmal-
eamide I) have very low activity for the atypical isoforms
(341, 536). Porcine relaxin was found to cause a concentra-
tion-dependent translocation of PKC-
to the cell mem-
brane in a number of cell lines: MCF-7 (human breast can-
cer), PHM1–31 (pregnant human myometrial), MMC
(mouse mesangial), and human monocytic (THP-1) cells
(382). PKC-
translocation was dependent on PI3K activa-
tion by relaxin, but independent of cAMP accumulation
(FIGURE 7A); furthermore, PI3K-dependent cAMP in-
creases mediated by relaxin were dependent on PKC
expression (382). In MCF-7 cells, relaxin stimulation ac-
tivates PI3K and causes translocation of PKC-
without
increasing cAMP accumulation; however, subsequent
transfection of AC5, but not AC2 or AC4, into these cells
produced a cAMP response to relaxin (381). Thus
RXFP1 activates PI3K, precipitating the translocation of
PKC
to the cell membrane, whereby the protein stimu-
lates AC5 (FIGURE 7A).
This pathway has also been demonstrated in HEK293 cells
transiently or stably expressing RXFP1 (198, 204), and was
found to be downstream of G
and G
␤␥
subunits. Experi-
ments utilizing G
i/o
mutants that are insensitive to ADP
ribosylation by pertussis toxin (PTX) revealed initial cou-
pling of RXFP1 to inhibitory G
OB
(in addition to G
s
) and
identified G
i3
as the mediator of the G
␤␥
-PI3K-PKC-
-
AC5 pathway. Interestingly, a PTX-sensitive stimulation of
cAMP had been previously demonstrated in an endogenous
setting: in rat left atria, relaxin induced a small increase in
cAMP accumulation that was reduced by PTX pretreat-
ment, and PTX pretreatment also reduced the inotropic and
chronotropic responses to relaxin (290), indicating the
physiological relevance of this signaling pathway. This
pathway has recently been shown to be of greater impor-
tance in the failing heart (typically characterized by in-
creased G
i/o
expression) (66, 150), allowing preservation
of the positive inotropic effects of relaxin in humans (127).
Site-directed mutagenesis identified the final 10 amino acids
of the RXFP1 COOH terminus, and in particular R752 as
essential residues for activation of the G
i3
pathway (FIG-
URE 3) (204). The activation of this pathway is also depen-
dent on the presence of lipid-rich membrane domains, sug-
gesting a degree of specific compartmentalization of the
RXFP1-stimulated cAMP response (204). Furthermore,
GTP
S-immunoprecipitation assays identified activation of
G
i3
immediately following stimulation of RXFP1 by re-
laxin, thus suggesting the delay in activation of the G
i3
-
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
433Physiol Rev VOL 93 JANUARY 2013 www.prv.org
G
␤␥
-PI3K-PKC
-AC5 pathway is downstream of the G
protein itself (204). In HEK293 cells, only activation of G
s
and G
OB
-dependent cAMP signaling pathways increases
CRE-mediated gene transcription, whereas G
i3
-mediated
signaling appears to selectively regulate NF
B-dependent gene
transcription (FIGURE 7A)(197). This observation again sug-
gests compartmentalization of relaxin-RXFP1 signaling
events and infers that distinct physiological outcomes could be
anticipated downstream of different cAMP signaling path-
ways (197).
Recently, a constitutive RXFP1-dependent cAMP response
has been identified using FRET-based cAMP biosensors in
single rat cardiac fibroblasts, HeLa cells, and HEK293 cells
expressing RXFP1 (201). The response is dependent on a
protein complex, or signalosome, centered upon the relaxin
receptor, and the signalosome is highly sensitive to attomo-
lar concentrations of relaxin. Importantly, this may provide
the basis for cellular responses to low levels of relaxin pres-
ent in the circulation. The signalosome consists of RXFP1
that is scaffolded to AC2 by AKAP79, facilitating efficient
activation of the AC by G
s
and G
␤␥
subunits. The cAMP
produced is tightly regulated by the activity of protein ki-
nase A-activated PDE4D3, that is in turn scaffolded to the
receptor COOH terminus (specifically requiring S704) by
-arrestin 2 (FIGURE 8A)(201). The stimulatory (AKAP79
and AC2) and regulatory (
-arrestin 2, PKA, and PDE4D3)
arms of the signalosome are both spatially and functionally
distinct. Knockdown of AKAP79 does not affect interaction
of regulatory components with the receptor, and knock-
down of
-arrestin 2 does not influence interactions be-
tween RXFP1 and the stimulatory components. Targeted
protein knockdown or overexpression of dominant nega-
tive mutants (201) demonstrates that the complex is iso-
form specific (i.e., interacting with PDE4D3 but not
PDE4D5) (234) and that assembly of the regulatory com-
ponents is dependent on constitutive association between
the receptor and
-arrestin 2, but not
-arrestin 1 (FIGURE
8A)(120). cAMP production by the RXFP1-signalosome is
dependent on constitutive association between helix 8 of
the receptor and AKAP79, but not gravin (AKAP250) or
AKAP149 (21, 122, 484). Importantly, this signaling mech-
anism is distinct from those activated by higher concentra-
tions of relaxin, and the complex dissociates following ac-
RXFP3
D
Relaxin-3 signalling bias
PLC
PI3K
AP1 Transcription
Src
Shc
SOS Ras
Raf
MEK1/2
ERK1/2
NFκB Transcription
PKC IκB
Relaxin-3
β
γ
Internalization
p38
JNK
MKK4/7
RXFP3
E
Relaxin signalling bias
PLC
AP1 Transcription
Ras
Raf
MEK1/2
ERK1/2
PKC
Relaxin
βγ
p38 JNK
MKK4/7
A
PDE4D3
cAMP
β
γ
PKA
AKAP79
Basal
B
PDE4D3
ATP
cAMP
βγ
PKA
AKAP79
Sub-picomolar relaxin
Signalosome activation
CNanomolar relaxin
Signalosome dissociation
PDE4D3
PKA
AKAP79
Activation of cannonical
cAMP signalling pathways
Gαs
β-arrestin 2
S704
AMP
ATP
RXFP1 AC2 RXFP1 AC2
Gαs
β-arrestin 2
S704
AMP
GαoB Gαs
Gαi3
RXFP1 AC2
β-arrestin 2
AC
GαoB
Gαi2
GαoBGαi2
AC
FIGURE 8. Atypical signaling mechanisms utilized by RXFP1 and RXFP3. Signaling complexes termed
signalosomes are formed between RXFP1, AKAP79, AC2, PKA, PDE4D3, and
-arrestin-2 (A) (196, 201).
These complexes are exquisitely sensitive and respond to attomolar concentrations of relaxin with increases in
cAMP (B). This signaling mechanism is distinct from those utilized by higher concentrations of relaxin, and
exposure of the receptor to nanomolar concentrations causes dissociation of the signalosome (C) and cAMP
production by the canonical pathways (see FIGURE 7). RXFP3 displays signaling bias with respect to the
cognate peptide human relaxin-3 and human relaxin-2 (545). Interaction of relaxin-3 with RXFP3 leads to
inhibition of adenylyl cyclase and activation of MAP kinases, NF
B and AP-1 transcription (D), whereas human
relaxin-2 causes only a subset of these pathways to be activated (E).
BATHGATE ET AL.
434 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
tivation of RXFP1 with nanomolar concentrations of pep-
tide that increase cAMP by previously described mechanisms
(FIGURES 7AAND 8C) (198). The high degree of sensitivity
exhibited by the RXFP1 signalosome has been demon-
strated previously in a few other physiological systems, in-
cluding suppression of pro-inflammatory cytokine produc-
tion by interleukin-15 (6), proliferation of helper T-cells by
interleukin-1 (398), the effects of neuropeptides and neuro-
steroids in nociception (447, 537), and the long-term effects
of transforming growth factor-
on basal FSH levels (575).
However, although signalosome formation and activity has
also been demonstrated in primary cells endogenously ex-
pressing RXFP1, the exact physiological role for the mech-
anism has yet to be determined.
Relaxin-induced cAMP accumulation may also occur via a
G protein-independent mechanism, and in some cells in-
creased cAMP accumulation may be downstream of a ty-
rosine kinase. In THP-1 cells and cultures of primary
human myometrial or endometrial stromal cells, porcine
relaxin induced an increase in cAMP accumulation that
was blocked by inhibition of tyrosine kinase activity (13,
38, 216, 302). This same response was potentiated by the
phosphotyrosine phosphatase inhibitors [bpV(phen) and
mpV(pic)] that mimic the effect of tyrosine kinase activa-
tion (38). Indeed, in studies predating receptor identifica-
tion in human lower uterine segment fibroblasts, relaxin
stimulation resulted in tyrosine phosphorylation of cellular
extracts, with no effect on cAMP accumulation (407). Re-
laxin-mediated increases in cAMP accumulation in this
context are thought to occur via a tyrosine kinase-mediated
inhibition of a PDE, thereby preventing cAMP hydrolysis
and thus increasing cAMP levels. Interestingly, studies have
also shown that the same tyrosine kinase inhibitors did not
affect relaxin-stimulated cAMP accumulation in HEK293
cells expressing RXFP1 (13), emphasizing the variation in
cellular responses that can be observed between different
cell types. In those cells utilizing the tyrosine kinase path-
way however, there was also some degree of cAMP inhibi-
tion in the presence of the PI3K inhibitor, LY294002 (13,
216), and evidence of a negative feedback loop involving
PKA (13).
A number of other studies in an endogenous setting have
also demonstrated activation of PKA downstream of
cAMP. Increased cAMP accumulation by relaxin results in
PKA activation in a human myometrial cell line (124). In
this case, PKA (via its interaction with AKAP79) can inhibit
the phosphoinositide turnover that is stimulated by oxyto-
cin and is required for smooth muscle contraction (124,
592). In the same cell line, human relaxin can also stimulate
aCa
2
-sensitive K
channel independently of increases in
Ca
2
; a PKA inhibitor, Rp-cAMPS, prevented the effects of
relaxin, and the stimulation was mimicked by the addition
of PKA
-catalytic subunits to the preparation (364). Fur-
thermore, the inhibition of oxytocin-induced events by re-
laxin is attributed to activation of PKA, and is prevented
following the use of a PKA inhibitor (16, 247). Thus relaxin
stimulation of RXFP1 absolutely results in the activation of
cAMP-PKA pathways, although the precise mechanism
whereby RXFP1 initiates cAMP accumulation appears to
vary somewhat with cell type.
2. Activation of RXFP1 causes increased
phosphorylation of mitogen-activated protein
kinases
In addition to increased cAMP accumulation, a number of
cell types that express RXFP1 including human endometrial
stromal cells (586), THP-1 cells and primary cultures of
human coronary artery cells, pulmonary artery smooth
muscle cells, renal myofibroblasts (370), and fibrochondro-
cytes (4) respond to relaxin with a rapid activation of
p42/44 MAPK (extracellular signal regulated kinase 1/2,
ERK1/2) (FIGURE 9).
In normal human endometrial cells (NHE cells), relaxin
stimulation causes rapid and transient phosphorylation of
ERK1/2, with a peak response between 5–10 min (586).
The same time course profile following relaxin stimulation
was also recorded for phosphorylation of MEK and CREB,
but in these cells there was no effect of relaxin treatment on
Akt or JNK phosphorylation; furthermore, an inibitor of
MEK abolished phosphorylation of ERK1/2 in response to
relaxin, suggesting that MEK is activated upstream of
ERK1/2 (586). Increased phosphorylation of ERK1/2 fol-
lowing relaxin stimulation was also observed in THP-1 cells
and cultures of human coronary artery and pulmonary ar-
tery smooth muscle cells and the MEK-ERK1/2 pathway
was also found to control relaxin-stimulated transcription
of VEGF in these cells (FIGURE 9) (586).
In contrast to the brief increase in ERK1/2 phosphorylation
described above, relaxin stimulation caused a more prolonged
increase in ERK1/2 phosphorylation in HeLa cells and pri-
mary human umbilical vein endothelial cells (HUVECs) (129).
Furthermore, in HeLa, EAhy926 (an endothelial cell line),
HT-29 (a colonic cell line), and primary fibrochondrocyte
cells, relaxin increased the phosphorylation of both ERK1/2
and Akt after 30 min (FIGURE 9) (4, 128). In primary fibro-
chondrocytes, it has also been shown that in addition to
ERK1/2 and Akt, treatment with relaxin also activates PI-
3-kinase, PKC-
,NF
B, c-fos, and Elk-1 all of which influ-
ence the expression of MMP-9 (4) (see sect. VIIA5). Simi-
larly, relaxin induces a sustained increase in ERK1/2 phos-
phorylation in rat renal myofibroblasts, and this was
potentiated by inhibition of G
i/o
by PTX, suggesting that
phosphorylation of ERK1/2 may be downstream of G pro-
tein coupling (370). In contrast, in human vascular smooth
muscle cells (hVSMC), relaxin stimulation did not affect
ERK1/2 phosphorylation but instead increased the phos-
phorylation of p38 MAPK (129). Thus, although relaxin
increases phosphorylation of a number of kinases in multi-
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
435Physiol Rev VOL 93 JANUARY 2013 www.prv.org
ple cell types, the precise isoform and mode of activation
appears to vary, and the physiological consequences of ac-
tivation of these pathways are as yet unclear.
3. Role of NO signaling in RXFP1-mediated
responses
There is a large volume of evidence that shows relaxin treat-
ment increases NO synthesis both acutely and chronically,
although the exact mechanism and whether this is directly
or indirectly linked to RXFP1 is still unclear (104, 385).
Several physiological effects of relaxin within the cardiovas-
cular system are mediated by NO, including inhibition of
lipopolysaccharide (LPS)-induced neutrophil adhesion in
coronary endothelial cells (386), inhibition of the activation
of neutrophils by proinflammatory agents via increased in-
ducible NO synthase (36) expression (345), increased cor-
onary blood flow in rat and guinea pig hearts (36), and the
increased renal vasodilation and hyperfiltration in rats via
the ET
B
receptor (FIGURE 9) (113). Relaxin also increases
CREB
VEGF
ERK
Gαi3 Gαo Gαi3 GαoB
GαS
ETB
ET-1
RXFP1
Relaxin
ET-1-32
VSM
bigET-1
AKT
NO
PDE
PKCζ
Adenylate
cyclase
PKA
cAMP
ATP AMP
IκB
NFκB
eNOS
nNOS
iNOS
Nucleus
Altered gene transcription
MMPs
2, 9
βγ βγ
PI3K
FIGURE 9. Vasodilator mechanisms activated by relaxin acting at RXFP1 receptors. Relaxin induces vaso-
dilatation by activation of multiple mechanisms. Vasodilation occurs both acutely and chronically, and the action
of relaxin differs in different blood vessels and in humans even with medication being administered (171). Acute
actions that likely influence vascular tone include cAMP generation and increases in NO generation (104, 385).
Chronic actions include the induction of eNOS, nNOS, and iNOS that in addition to the direct effects of NO cause
increased expression of MMPs that cleave big ET-1 to release ET(1–32) that acts on ET
B
receptors to cause
NO-dependent vasorelaxation (265).
BATHGATE ET AL.
436 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
the activity and expression of the three types of NOS: en-
dothelial NOS (eNOS; NOS III) (19, 20, 27, 28), inducible
NOS (iNOS; NOS II) (20, 28, 30, 31, 89, 154, 345, 346,
427), and neuronal NOS (nNOS; NOS I) (19) (FIGURE 9).
Furthermore, activation of NOS by relaxin results in down-
stream increases in cGMP, due to production of NO and
activation of guanylyl cyclase (30–32, 346). Based on acti-
vation of NOS and cGMP, there are a number of pathways
by which relaxin could increase NO production. First,
eNOS activity can be increased by Akt phosphorylation;
thus coupling of RXFP1 to G
i3
and the associated release
of G
␤␥
subunits can activate PI3K, that could potentially
also activate Akt, allowing phosphorylation of eNOS at
S1172 (385) (FIGURE 9). Second, iNOS activity can be up-
regulated following stimulation of NF
B-controlled tran-
scription; thus coupling of RXFP1 to G
s
increases cAMP
and activates PKA, which can then potentially phosphory-
late and inactivate I
B, thereby stimulating NF
B-con-
trolled transcription and increasing iNOS-mediated NO
production (31, 154). Indeed, the increase in iNOS expres-
sion mediated by relaxin in HUVECs (427) and rat coro-
nary endothelial cells (154) was abolished by NF
B inhib-
itors (FIGURE 9). More evidence supporting this possibility
comes from studies that demonstrate stimulation of NF
B
transcription by relaxin over short time periods (129, 130,
224), suggesting that this pathway may represent a mecha-
nism for more transient increases in NO (FIGURE 9). Finally,
the pathway whereby relaxin stimulation of RXFP1 could
upregulate eNOS in the kidney is better defined: it is hy-
pothesized that this physiological effect of relaxin is due to
increased activity of matrix metalloproteinase (MMP) 2
(445) or MMP9 (215, 224), which can process big endothe-
lin (ET) to ET
1–32
,ET
1–32
then activates the ET
B
receptor,
increasing the activity of eNOS, and thus producing NO
(FIGURE 9) (263).
4. RXFP1 signaling influences connective tissue
metabolism
Relaxin can modulate connective tissue metabolism at a
number of levels, including the inhibition of profibrotic fac-
tors (such as TGF-
), the inhibition of fibroblast prolifera-
tion and differentiation, and the activation of MMP-medi-
ated extracellular matrix degradation. Identification of the
signaling pathways that lead to these end points are now
gaining attention, and greater knowledge of these processes
should allow more targeted efforts in relaxin-related drug
discovery.
TGF-
is one of the most important cytokines in the pro-
gression of fibrosis; binding of TGF-
to its type II receptor
leads to formation of a receptor complex with the type I
receptor, receptor phosphorylation, activation of a receptor
(R)-Smad (Smad1, 2, 3, 5 or 8), and cotransport of R-Smad
with a common (Co)-Smad (Smad4) into the nucleus to
control gene expression (reviewed in Ref. 347). The antifi-
brotic effects of relaxin are thought to result from inhibition
of the action of TGF-
(FIGURE 10) (349, 443, 539, 542). In
human renal fibroblasts, TGF-
increases the expression of
-SMA (a marker of fibroblast differentiation), type I col-
lagen, and fibronectin, and these effects are reversed by
relaxin. The inhibitory effects of relaxin are due to inhibi-
tion of Smad2 (an R-Smad) (215, 370) but relaxin treat-
ment also decreases TGF-
-induced phosphorylation of
Smad2 (but not Smad3), and knockdown of Smad2 alone
had an effect similar to treating the cells with relaxin (215).
Subsequent studies in rat renal myofibroblasts confirmed
inhibition of Smad2 phosphorylation by relaxin and
identified the upstream mediators of this inhibition:
nNOS-NO-cGMP (370). Thus relaxin inhibits TGF-
and thus fibroblast differentiation by a NO-dependent
pathway (FIGURE 10).
Relaxin also affects connective tissue metabolism by in-
creasing the production of MMPs including MMP1 in hu-
man lung fibroblasts (542), human dermal fibroblasts (539)
and rat renal cortical fibroblasts (349), MMP2 in rat car-
diac fibroblasts (443) and renal fibroblast cell lines (215),
MMP9 in renal fibroblast cell lines (215) and THP-1 cells
(224), and MMP13 in rat hepatic stellate cells (FIGURE 10)
(55). Increased MMP9 activity in THP-1 cells (that endog-
enously express relaxin receptors) was due to changes in
transcriptional regulation of the MMP gene mediated by
NF
B, as inhibition of NF
B reduced the relaxin-dependent
increase in MMP9 expression and activity (FIGURE 10)
(224). This may suggest an interesting link between the
effects of relaxin upon MMPs and NO, both of which ap-
pear to be mediated by increased NF
B transcription. To
complement the increased production of MMPs in response
to relaxin, the peptide can also decrease the expression of
tissue inhibitors of MMPs (TIMPs) in human dermal fibro-
blasts (539), with specific inhibition of TIMP1 and TIMP2
in cultured rat hepatic stellate cells (55, 564).
A recent report indicates that relaxin increases the activity
of the peroxisome proliferator-activated receptor (PPAR)-
(482), a nuclear receptor that heterodimerizes with the ret-
inoid X receptor to increase the transcription of target genes
(reviewed in Ref. 153). PPAR
has similar inhibitory effects
on TGF-
signaling and similar end-point physiological ef-
fects upon fibrosis to relaxin itself (482). Thus this obser-
vation may provide another missing link in the signaling
pathways associated with connective tissue metabolism
that are activated by relaxin. However, any physiological
relevance of this activation, and indeed the relevance of this
to relaxin signaling pathways, remains to be shown.
B. Signaling Pathways Activated by RXFP2
Although relaxin and INSL3 are closely related peptides
and RXFP1 and RXFP2 have similar structures, the intra-
cellular signaling pathways that are initiated following
INSL3 stimulation of RXFP2 are less elaborate. INSL3 (or
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
437Physiol Rev VOL 93 JANUARY 2013 www.prv.org
relaxin) stimulation of RXFP2 expressed in HEK293T cells
results in coupling to G
s
to increase cAMP accumulation,
and to G
oB
to negatively modulate increases in cAMP
(FIGURE 7B)(198, 301). RXFP2 signaling therefore closely
resembles the first stage of RXFP1 signaling in response to
relaxin (198). There is also evidence for an inhibitory effect
of G
␤␥
subunits upon AC; the G
␤␥
subunits are likely de-
rived from G
oB
, as although removal of G
␤␥
using
ARK-ct or inhibition of G
i/o
with PTX both increased
cAMP the effects were not additive (198). Unlike RXFP1,
for RXFP2 there is no evidence for constitutive activity or a
high sensitivity response to INSL3 (201), and in addition,
RXFP2 does not activate the G
i3
-cAMP signaling pathway
that is unique to RXFP1 (198, 204). Downstream of cAMP
production, activation of RXFP2 (by either relaxin or
INSL3) induces increased CRE-dependent gene transcrip-
tion (FIGURE 7B)(197) in a manner similar to the relaxin
response through RXFP1.
The effect of both G
s
and G
oB
upon cAMP accumulation
has also been shown in cells that endogenously express
RXFP2. INSL3 stimulation of RXFP2 in rat gubernacular
cells (301) and in a human osteoblast cell line (MG-63)
(162) leads to increased cAMP accumulation, probably by a
G
s
-mediated interaction with AC, similar to the response
in HEK293T cells expressing RXFP2. However, in con-
ERK
Gαi3 GαoB
GαS
TGFβ
RXFP1
Relaxin
AKT
NO
PDE
PKCζ
Adenylate
cyclase
PKA
cAMP
ATP AMP
IκB
NFκB
eNOS
nNOS
iNOS
Nucleus
Altered gene transcription
MMPs
1, 2, 9, 13
Fibrosis
Collagen
αSMA
βγ
PI3K
P
P
P
P
pSmad2
FIGURE 10. The effects of relaxin acting on RXFP1 to influence connective tissue metabolism. The antifi-
brotic actions of relaxin are mediated through increases of expression of eNOS, nNOS, and iNOS and increased
synthesis of NO. This in turn increases expression of MMPs to reduce fibrosis and also inhibits pSmad2, a
mediator involved in the profibrotic effects of TGF-
.
BATHGATE ET AL.
438 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
trast, primary cultures of testicular germ cells and oocytes
respond to INSL3 activation of RXFP2 with a PTX-sensi-
tive inhibition of cAMP accumulation (276), consistent
with RXFP2 coupling to G
oB
. Thus, as for RXFP1, the net
signaling outcome of RXFP2 stimulation will depend on
which signaling components (especially G protein isoforms)
are expressed in a particular cellular background.
C. Signaling Pathways Activated Following
Stimulation of RXFP3
Activation of RXFP3, by human relaxin-3, causes PTX-
sensitive inhibition of forskolin-stimulated cAMP accumu-
lation, suggesting that the receptor is coupled to inhibitory
G
i/o
proteins. RXFP3 activation by human relaxin-3 also
induces phosphorylation of ERK1/2 and other MAPKs
through G
i/o
via either a PI3K-dependent or PKC-depen-
dent mechanism (FIGURE 7C)(545, 546). Although initial
studies on RXFP3 suggested human relaxin-3 or the
-chain of human relaxin-3 were the only relaxin family
peptides that could compete for
125
I-human relaxin-3 bind-
ing or inhibit AC activity (314), more recent work has dem-
onstrated that human relaxin-2 also has agonist actions at
RXFP3 albeit activating different signaling pathways (see
below) (545). Thus human relaxin-2 acts as a biased ligand
relative to human relaxin-3 at RXFP3, which may be sig-
nificant for the development of compounds with novel
modes of action and specificity at this receptor (FIGURE 8, D
AND E).
1. G protein coupling and second messenger
signaling at RXFP3
Whole cell functional responses downstream of RXFP3 ac-
tivation have been examined using the cytosensor micro-
physiometer that detects changes in whole cell metabolism
by measuring fluctuations in extracellular pH, that occur
due to the extrusion of H
ions as a by-product of cell
signaling. In CHO-K1 cells stably expressing the human
RXFP3 receptor, this approach provided more evidence
that RXFP3 signaling occurred mainly through the engage-
ment of G
i/o
proteins, as pretreatment with PTX blocked
human relaxin-3-stimulated increases in the extracellular
acidification response (543). This was consistent with pre-
vious observations that human relaxin-3 inhibited forsko-
lin-stimulated cAMP accumulation, presumably by engage-
ment of G
i/o
proteins (314). The use of inhibitors of PI3K
and MEK 1/2 also indicated that signaling pathways down-
stream of these two kinases were activated by human relax-
in-3 (see sect. VIB3). In addition, a second generalized cell
signaling screen using eight different reporter gene con-
structs in CHO-K1 cells (transiently expressing human
RXFP3 receptors) showed that following stimulation with
human relaxin-3, RXFP3 activates signaling upstream of
activator protein 1 (AP-1) promoters (see sect. VIB4) and
NF
B promoters (see sect. VIB5), but not serum response
element, CRE, heat shock element, nuclear factor of acti-
vated T-cells element, E-box DNA binding element, or glu-
cocorticoid response element (545).
2. RXFP3 is coupled to the inhibition of cAMP
accumulation in recombinant and endogenously
expressing systems
Following stimulation with human relaxin-3, RXFP3 inhib-
its forskolin-stimulated cAMP accumulation in CHO-K1 or
HEK293 cells stably expressing RXFP3 receptors and in
mouse SN56 cells that endogenously express the receptor
(FIGURE 7C)(314, 546). Human relaxin-3-mediated inhibi-
tion of AC is completely prevented in cells pretreated with
PTX, suggesting that this response is G
i/o
dependent. In
CHO-K1 cells transiently transfected with PTX-insensitive
(C351I mutation) variants of G
i/o
proteins, and treated
with PTX (to remove the influence of endogenous G
i/o
proteins), G
i2
was the major G protein involved in the
inhibition of forskolin-stimulated cAMP accumulation,
whereas in HEK293 cells, G
i3
,G
oB
, and G
oA
were all
important (544). Together, these observations indicate that
although the same downstream signal output may be ob-
served, the signaling partners involved may vary with the
different cellular context.
Forskolin-stimulated cAMP accumulation is also inhibited
in RXFP3-expressing cells by human relaxin-3 B-chain pep-
tides or to a lesser degree by human relaxin-2 and porcine
relaxin in the CHO-K1, HEK293, and SN56 cell back-
grounds (314, 545, 546). These responses likely also occur
via RXFP3 engagement of G
i/o
proteins, although this re-
mains to be determined.
3. Activation of RXFP3 increases phosphorylation
of ERK1/2
Human relaxin-3 activation of RXFP3 causes a rapid and
transient increase in ERK1/2 phosphorylation (peak re-
sponse 2–5 min) in CHO and HEK293 cells, stably express-
ing human RXFP3 (FIGURE 7C)(545, 546). Examination of
a variety of relaxin family peptides revealed that in addition
to human relaxin-3, the human relaxin-3 B-chain dimer
also activates ERK1/2 albeit with low potency and efficacy
(546). There was also a weak response to human relaxin-2
but no detectable response to porcine relaxin or INSL3
(545). Pretreatment with PTX caused an 90% inhibition
of human relaxin-3 stimulated ERK1/2 phosphorylation in
CHO-RXFP3 cells, whereas in HEK-RXFP3 cells the signal
was completely abolished (546). This suggests that the cou-
pling of RXFP3 to PTX-sensitive G
i/o
proteins mediates
ERK1/2 phosphorylation, in both recombinant and endog-
enous systems (546).
RXFP3-mediated ERK1/2 phosphorylation occurs via two
parallel pathways, both of which are downstream of G
i/o
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
439Physiol Rev VOL 93 JANUARY 2013 www.prv.org
in both CHO-K1 and HEK293 cells stably expressing hu-
man RXFP3 receptors, and in SN56 cells endogenously ex-
pressing mouse RXFP3 receptors. The first pathway is de-
pendent on PI3K activation (50% blockade of human
relaxin-3 stimulated ERK1/2 phosphorylation by the PI3K
inhibitors LY294002 or wortmannin), while the other
pathway requires PKC (partial blockade of ERK1/2 phos-
phorylation by both general and isoform-selective PKC in-
hibitors) (FIGURE 7C)(546). Interestingly, many MAPKs,
including ERK1/2, are implicated in central feeding re-
sponses in rats (374, 448, 468), which may suggest that
activation of ERK1/2 by RXFP3 is a physiologically rele-
vant signaling pathway, although direct links between
RXFP3, ERK1/2 activation, and increases in feeding remain
to be shown.
4. RXFP3 activates AP-1-linked reporter genes
Since multiple MAPK signaling pathways [p38 MAPK
(436), JNK (116), and ERK1/2 (423, 558)] converge on
AP-1 elements to increase gene transcription, and are likely
involved in activation of the AP-1-linked reporter genes, it
was postulated that RXFP3 may also couple to other
MAPK pathways (FIGURE 8, DAND E) (see sect. VIB3).
Indeed, human relaxin-3-mediated AP-1 reporter gene acti-
vation was completely blocked in CHO-RXFP3 cells pre-
treated with the p38 MAPK inhibitor (RWJ67657),
whereas a MEK inhibitor (PD98059) or a JNK inhibitor
(SP600125) shifted the human relaxin-3 concentration-re-
sponse curve to the right. In contrast, in HEK-RXFP3 cells,
JNK inhibition abolished human relaxin-3-stimulated AP-1
reporter activation, whereas p38 MAPK or MEK inhibition
shifted concentration-response curves to the right. In SN56
cells (as in CHO-K1 cells), p38 MAPK inhibition com-
pletely abolished AP-1 reporter gene activation and JNK
and MEK inhibition partially blocked AP-1 activation
(545). This suggests that all three MAPKs are involved in
human relaxin-3-mediated AP-1 activation, and that the
hierarchy of the different signaling pathways varies with the
cell background.
Pretreating the cells with PTX blocked the human relaxin-
3-stimulated AP-1 reporter activation in SN56 cells but not
in CHO-RXFP3 and HEK-RXFP3 cells (545). This suggests
that while AP-1 reporter gene activation was downstream
of G
i/o
in a mouse-derived cell line, this same pathway was
activated by a G
i/o
-independent pathway downstream of
the human RXFP3 receptor (545).
These MAPK signaling pathways may have physiological
significance since in forced swim tests in rats, there are
dramatic increases in phosphorylated MEK1/2, ERK1/2,
and JNK1/2/3 immediately following the swim test (468),
suggesting that both ERK1/2 and JNK1/2/3 are important
in mediating central stress responses. In addition, human
relaxin-3 mRNA was increased in the nucleus incertus
(516), which may correlate with the increase in MAPK acti-
vation, although the direct links between human relaxin-3,
RXFP3, MAPK phosphorylation, and stress responses remain
to be demonstrated in brain.
5. RXFP3 is coupled to NF
B signaling
Activation of RXFP3 by human relaxin-3 increased NF
B
reporter gene activation in CHO and HEK293 cells tran-
siently expressing human RXFP3, and also in SN56 cells
endogenously expressing mouse RXFP3, and this activation
was blocked by PTX pretreatment (545). This again sug-
gests that NF
B activation occurs downstream of G
i/o
pro-
tein engagement by the receptor; however, the precise
mechanisms and physiological significance of this pathway
remain to be determined (FIGURE 8D).
6. Ligand-directed signaling bias at RXFP3
Ligand-directed signaling bias is gaining increasing cre-
dence in GPCR pharmacology. It describes the selective
stabilization of particular receptor confirmations by ligands
resulting in selective activation of downstream signal trans-
duction pathways (22, 151, 277). There is now evidence
that several relaxin peptides interact with RXFP3 to acti-
vate distinct signaling profiles through different, although
sometimes overlapping pathways (FIGURE 8, DAND E). Al-
though the original description of RXFP3 indicated that
this receptor showed a high selectivity for human relaxin-3
in both binding and AC inhibition assays (314), cross-reac-
tivity with other relaxin peptides was not explored over a
wider range of signal transduction pathways. However,
monitoring the whole cell effects of relaxin family peptides,
using the cytosensor microphysiometer and reporter gene
assays, provided the first indication that other relaxin fam-
ily peptides may elicit responses through RXFP3. Examina-
tion of the metabolic responses recorded by microphysiom-
etry indicated that human relaxin-2 caused a small change
in the extracellular acidification rate in CHO cells stably
expressing RXFP3 (543). In addition, when human relax-
in-2, porcine relaxin, and human INSL3 were tested in the
reporter gene paradigm, human relaxin-2 was identified to
have greater potency and efficacy at the AP-1 reporter gene
pathway than the cognate ligand human relaxin-3; this is a
hallmark of ligand-biased signaling, and suggests that this
phenomenon also occurred downstream of the RXFP3 re-
ceptor (FIGURE 8, DAND E)(545).
Reexamination of cAMP signaling in three distinct cellular
backgrounds (CHO, HEK293, and SN56) revealed strong
inhibition of forskolin-stimulated cAMP accumulation by
human relaxin-3. However, human relaxin-2, porcine re-
laxin, and human INSL3 also inhibited the forskolin-stim-
ulated cAMP accumulation albeit with much lower potency
than human relaxin-3 (545). Interestingly, some species-
specific inhibition of forskolin-stimulated cAMP accumula-
BATHGATE ET AL.
440 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
tion was noted, with INSL3 weakly inhibiting cAMP accu-
mulation downstream of the human RXFP3 receptor, but
not downstream of the mouse RXFP3 receptor (545). This
is in contrast to a previous study, which reported no inhi-
bition of forskolin-stimulated cAMP accumulation by ei-
ther porcine relaxin or human INSL3 (314). Since the sen-
sitivity of inhibitory cAMP assays is highly dependent on
both the degree of activation of AC by forskolin and the
time of stimulation, the differences observed between the
two studies most likely result from different experimental
paradigms.
Addition of human relaxin-2, porcine relaxin, and INSL3 to
RXFP3-expressing cells caused AP-1 reporter gene activa-
tion in a concentration-dependent and cell-type-specific
manner (545). AP-1 was activated in CHO-RXFP3 and
HEK-RXFP3 cells with an order of potency of human re-
laxin-2 human relaxin-3 porcine relaxin. Human re-
laxin-2 also activated AP-1 in SN56 cells with a similar
pattern to that in CHO-RXFP3 and HEK-RXFP3 cells, al-
though the pEC
50
was shifted to the right (545). Given the
differences observed between different species relaxins at
RXFP1 (see sect. IIC1), this shift in potency order may be
reversed with the use of mouse relaxin instead of human
relaxin-2. In CHO-RXFP3 cells, human relaxin-2-mediated
AP-1 reporter gene activation was strongly inhibited by the
p38 MAPK inhibitor (RWJ67657) or the JNK inhibitor
(SP600125); however, the MEK inhibitor (PD98059)
shifted the human relaxin-2 concentration-response curve
to the right without decreasing the maximum response, im-
plicating p38 MAPK and JNK as the major MAPKs in-
volved in mediating this response downstream of human
RXFP3. Contrastingly, when the same response was exam-
ined in HEK293 cells expressing human RXFP3, human
relaxin-2-stimulated AP-1 reporter gene activation was
most strongly influenced by p38 MAPK or MEK inhibition
and was not affected by JNK inhibition, suggesting that p38
MAPK and ERK were not the major MAPKs involved in
mediating this response downstream of human RXFP3. In
the SN56 cell line, expressing mouse RXFP3, the p38
MAPK, JNK, and MEK inhibitors all equally blocked the
human relaxin-2-stimulated AP-1 reporter gene activation,
suggesting that all three kinases were equally important in
this cellular context (545). Together these results suggest
that the cellular background, in addition to the species ho-
molog of the receptor, play important roles in the pattern of
activation of AP-1 reporter genes. Direct measurement of
ERK1/2, p38 MAPK, and JNK phosphorylation following
stimulation with the relaxin family peptides has confirmed
the findings of these inhibitor-based studies. Treating
CHO-RXFP3 or HEK-RXFP3 cells with human relaxin-2
increased ERK1/2 phosphorylation in a rapid and transient
manner with a peak response at 2–5 min; however, this
response was not observed in the SN56 cell background
(545) (Kocan et al., unpublished data).
Pretreatment with PTX failed to block either human relax-
in-3-stimulated AP-1 activation in CHO-RXFP3 or HEK-
RXFP3 cells, or porcine relaxin-stimulated AP-1 reporter
gene activation in CHO-RXFP3 cells, suggesting that hu-
man relaxin-3 and porcine relaxin activate AP-1 reporter
genes by a G protein-independent mechanism, again sug-
gesting ligand-directed signaling bias (545).
Although G protein-dependent activation of ERK1/2, p38
MAPK, and JNK by human relaxin-2 and porcine relaxin
were suggested initially by inhibitor studies, later confirmed
by direct MAPK phosphorylation assays, the G
i/o
-inde-
pendent pathway remains to be identified for human relax-
in-3 and porcine relaxin downstream of the human RXFP3
receptor. All of the AP-1 reporter gene responses arising
following stimulation of RXFP3 in SN56 cells were blocked
by PTX, suggesting that different signaling effectors were
involved in mediating the AP-1 reporter gene activation
downstream of the mouse RXFP3 receptor (545).
D. Signaling Pathways Activated by RXFP4
Little information currently exists about signaling path-
ways activated downstream of RXFP4. Stimulation of cells
recombinantly expressing human RXFP4 by INSL5 or hu-
man relaxin-3 increases GTP
S binding and inhibits fors-
kolin-stimulated cAMP accumulation, suggesting that
RXFP4 like RXFP3 is G
i/o
coupled (FIGURE 7D)(313,
315). When RXFP4 was coexpressed with G
16
in HEK293
cells, both INSL5 and human relaxin-3 stimulated a strong
Ca
2
signal in a concentration-dependent manner (313,
315), which may suggest that RXFP4 can couple to addi-
tional G proteins; however, direct evidence for RXFP4-G
protein coupling is lacking.
E. Identification of G Proteins That Couple
to RXFP Receptors
1. G proteins that couple to RXFP1 and RXFP2
The specificity of RXFP1 and RXFP2 for particular G
i/o
isoforms was identified using PTX-insensitive G protein
mutants (198). PTX inactivates G
i/o
proteins by ADP-ri-
bosylation of C351 so that mutation of this residue to I
renders the isoform of interest insensitive to inactivation by
PTX. Sequential expression of the three G
i
and two G
o
isoforms in HEK293T cells stably expressing either RXFP1
or RXFP2 caused the restoration of the expected cAMP
signaling profiles (following PTX pretreatment) by expres-
sion of only G
oB
for RXFP2 and following expression of
both G
oB
and G
i3
for RXFP1. Activation of G
oB
and
G
i3
by RXFP1, but the selective activation of G
oB
by
RXFP2 was subsequently confirmed using [
35
S]GTP
S im-
munoprecipitation; importantly, activation of G
i3
was ev-
ident within 3 min of relaxin stimulation, suggesting that
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
441Physiol Rev VOL 93 JANUARY 2013 www.prv.org
the observed delay in activation of the G
i3
-cAMP pathway
occurs downstream of receptor coupling to this G protein
isoform (204). Studies using receptor mutants further dem-
onstrated that activation of the G
i3
-cAMP pathway in-
volves the final 10 amino acids of the RXFP1 COOH ter-
minus and requires R752; interestingly, these 10 amino ac-
ids are missing in RXFP2 (the COOH terminus is 10
residues shorter), perhaps explaining the inability of this
highly similar receptor to activate G
i3
(204).
Due to the classification of both RXFP1 and RXFP2 as
GPCRs, and their ability to increase cAMP in recombi-
nant receptor expression systems, it has been assumed
that the receptors couple to G
s
. The ability of PTX to
enhance the cAMP generated following stimulation of
RXFP2 or in the early phase (10 min) following stimula-
tion of RXFP1 tends to support this view. More direct evi-
dence has been generated using peptide fragments of the
third intracellular loop of RXFP1 (minimum length 615–
629; ICL3) and G
s
; addition of the RXFP1-ICL3 alone to
rat tissues (striatum, cardiac and skeletal muscle membrane
preparations) increased cAMP and “antagonized” the re-
sponse to relaxin. Furthermore, addition of peptide frag-
ments of G
s
“antagonized” both relaxin stimulation and
RXFP1-ICL3 peptide stimulation of rat tissues (475). A
functional interaction between RXFP1 and G
s
has also
been demonstrated using pharmacological inhibition; ap-
plication of NF449 to HEK293 cells expressing RXFP1
significantly decreased receptor-stimulated cAMP accumu-
lation, whereas a low concentration of cholera toxin (a G
s
activator) substantially enhanced the cAMP response to re-
laxin (201). Taken together, this suggests that G
s
(and prob-
ably G
oB
) couples to RXFP1 within the ICL3, while coupling
to G
i3
is dependent on the final 10 amino acids of the receptor
COOH terminus.
2. G proteins that couple to RXFP3 and RXFP4
Functional assays suggest that RXFP3 and RXFP4 are G
i/o
-
coupled receptors that can inhibit forskolin-stimulated
cAMP accumulation in cells expressing the receptors (313,
314, 545). G
i/o
coupling to RXFP3 is further supported by
studies in the cytosensor microphysiometer, where pretreat-
ment of CHO-RXFP3 cells with the G
i/o
protein inhibitor
PTX strongly inhibited the extracellular acidification rate
response of these cells, suggesting that most of the signaling
pathways activated by RXFP3 occur downstream of G
i/o
(543). RXFP3 and RXFP4 do not activate calcium signaling
and are therefore presumed not to couple to G
q
proteins
(313, 314). This pattern of G protein coupling is supported
by more direct approaches. To examine the G protein sub-
units important for RXFP3 signaling, CHO-RXFP3 and
HEK-RXFP3 cells were transfected with PTX-insensitive G
proteins (C351I mutation), treated with PTX and then
ERK1/2 activation measured to determine which G proteins
could recapitulate the functional response. PTX pretreat-
ment completely abrogated the ERK1/2 response in both
CHO-RXFP3 and HEK-RXFP3 cells. In CHO-RXFP3
cells, the ERK1/2 response was partially restored by trans-
fection of G
i2
or G
oB
but not by mutant G
i1
,G
i3
or
G
oA
. In HEK-RXFP3 cells, the pattern differed somewhat
in that signaling was partially restored by expression of
mutant G
i2
or G
oB
but also by mutant G
oA
. These dif-
ferences may relate to different colocalization of receptors
and G proteins in lipid rafts in particular cell types.
F. Trafficking of RXFP Receptors
1. Trafficking of RXFP1 and RXFP2 to the cell
surface
Asparagine (N)-linked glycosylation is an important
posttranslational modification for cell surface expression
of GPCRs (430). RXFP1 has six predicted N-linked glyco-
sylation sites within the NH
2
-terminal tail, leucine-rich re-
peat, and LDLa modules, and RXFP2 has five predicted
N-linked glycosylation sites within the same region (203).
Western blot analysis of a hemagglutanin (HA)-tagged
RXFP1 showed multiple receptor bands, with a high-mo-
lecular-mass (95 kDa) band, representing the mature, gly-
cosylated form of the receptor and a lower molecular mass
(80 kDa) band, representing the immature form of the re-
ceptor (280). Digestion of RXFP1 with endoglycosidase H
or PNGase F removes the N-linked glycosylation from the
receptor, and only the lower molecular weight band is then
observable by Western blot (280). Similar results have been
achieved using a FLAG-tagged RXFP1 receptor where mu-
tagenesis of each of the predicted N-linked glycosylation
sites demonstrates that all such sites are utilized in RXFP1
(572). Additionally, glycosylation within the leucine-rich
repeats was shown to be important for both cell-surface
localization of the receptor and full cAMP signaling effi-
cacy, but has no role in ligand binding (572). As yet, the
glycosylation status of RXFP2 has not been studied in de-
tail.
In addition to glycosylation, the LDLa modules of RXFP1
and RXFP2 are important for receptor trafficking from the
ER to the plasma membrane. Expression of mutant RXFP1
receptors at the cell surface was not changed (compared to
native RXFP1), by deletion of the LDLa module, or by
mutation of the conserved LDLa residues, even though
these receptors could no longer activate cAMP signaling in
the cells (231, 280, 456). In one study, mutation of the
putative N-linked glycosylation site (N36Q) decreased both
cAMP production and cell surface expression (decreased to
37% compared with native RXFP1), suggesting that glyco-
sylation of RXFP1 in the LDLa module plays a role in cell
surface expression. However, an independent study on the
same RXFP1 (N36Q) mutant showed only slight decreases
in cell surface expression that was actually correlated with
decreases in total cell expression. In RXFP2, mutation of
the conserved cysteine residue (C71Y) or the conserved as-
BATHGATE ET AL.
442 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
partic acid residue (D70Y) (which destabilizes the LDLa
module) (231) decreased the expression of receptors at the
cell surface and compromised cAMP signaling (compared
with native RXFP2), also suggesting that the LDLa module
is important in cell surface expression of RXFP2 (64). In
addition, swapping the LDLa modules of RXFP1 and RXFP2
to produce a chimeric receptor improved cell surface delivery
of RXFP1, compared with native RXFP1 (280), suggesting
that the LDLa module of RXFP2 is also important for delivery
of the receptor to the plasma membrane.
RXFP1 receptors also form functional homodimers within
the ER, which are maintained while the receptors are traf-
ficked from the ER to the plasma membrane (282, 510,
511). Together these observations suggest that RXFP1 ex-
pression at the cell surface is regulated at a number of stages
during receptor synthesis and maturation, to ensure deliv-
ery of matured, functional receptors to the cell surface.
2. Limited evidence for trafficking of RXFP3 and
RXFP4
There are two conserved putative N-glycosylation sites in
the NH
2
terminus of both RXFP3 and RXFP4 and an ad-
ditional nonconserved site in ECL3 of human RXFP4; how-
ever, there is currently no information as to whether these
sites are utilized or are important for receptor trafficking.
Cell surface expression of RXFP3 and RXFP4 has only been
demonstrated by whole cell radioligand binding assays
(213, 545).
G. Regulation of RXFP Receptors
1. RXFP1 and RXFP2 are weakly regulated by
phosphorylation and internalization
A consistent feature associated with biological effects
mediated by RXFP1 is sustained signaling (see sect. IXA),
which is unusual for a GPCR. This is also observed when
RXFP1 and RXFP2 are expressed in HEK293T cells
where cAMP levels remain elevated for up to 6 h after
ligand exposure (87). RXFP1 and RXFP2 also display a
lack of internalization following exposure to ligand when
expressed in HEK-293T cells and Cos-7 cells (FIGURE 11)
(87). Both receptors are only weakly phosphorylated and
internalized as measured by whole cell radioligand bind-
ing, and internalization is unaffected by overexpression
of G protein-regulated kinase (GRK) 2/3, suggesting that
any internalization is not regulated by these isoforms.
However, GRK4, -5, and -6 were not tested, and it re-
mains to be determined whether RXFP1 and RXFP2 in-
ternalization is regulated by other GRKs. Confocal im-
aging of GFP-tagged
-arrestin also failed to show cell
surface localization of
-arrestin in cells cotransfected
with either RXFP1 or RXFP2 (FIGURE 11) (87), suggest-
ing that the receptors do not recruit either of the nonvi-
sual
-arrestin isoforms. In primary human decidual cells
and HEK293T cells stably transfected with RXFP1, weak
internalization occurs upon agonist stimulation as mea-
sured by cell surface ELISA, although this may also re-
flect the constitutive activity of the receptor (281). Fur-
thermore, it was suggested that overexpression of
-ar-
restin 2 in HEK-RXFP1 cells further reduced the number
of cell surface receptors (281). Interestingly, a recent
study provided evidence for constitutive association be-
tween RXFP1 (but not RXFP2) and
-arrestin 2, which
controls the endogenous constitutive activity of the re-
ceptor by scaffolding PDE4D3 (201). Thus
-arrestin
may be involved in RXFP1 signaling but is not recruited
following receptor stimulation nor is it associated with
the internalization of either RXFP1 or RXFP2.
2. RXFP3 internalization in response to agonist
exposure
Internalization of RXFP3 occurs following 10 min of
stimulation with human relaxin-3 in CHO-K1, HEK293,
and SN56 cell backgrounds, as assessed by radioligand
internalization assays (70–90% of receptors internal-
ized) (545). RXFP3 internalization was also observed by
confocal microscopy of a GFP-tagged RXFP3 following
stimulation with human relaxin-3 but not human relaxin-2,
porcine relaxin, or INSL3 in CHO-RXFP3 or HEK-
RXFP3 cells (545). Studies using BRET to detect interac-
tions between RXFP3 and
-arrestins in CHO cells
showed that human relaxin-3 but not human relaxin-2
promotes RXFP3/
-arrestin interaction (FIGURE 12).
This suggests that treatment with the cognate ligand hu-
man relaxin-3 causes RXFP3 to undergo
-arrestin-de-
pendent internalization. Pretreatment of cells with PTX
prevented 50% of these interactions, suggesting that
they are partially mediated by G
i/o
(Kocan et al., unpub-
lished data). The detailed mechanism of RXFP3 internal-
ization, phosphorylation, and recycling/degradation all
remain to be determined. To date, there are no data on
the regulation of RXFP4.
VII. PHYSIOLOGICAL ROLES OF PEPTIDE-
RECEPTOR PAIRS
A. Relaxin-RXFP1
Although RLN1 was the first human relaxin gene to be
cloned (245) and is expressed in decidua, trophoblasts, and
prostate (208, 439), its physiological role is unclear. A na-
tive human relaxin-1 has not been isolated, although a syn-
thetic human relaxin-1 peptide based on the structure of
human relaxin-2 does have similar biological properties
and potency to human relaxin-2 at RXFP1 receptors (514).
The RLN1 gene is only found in humans and the great apes,
but in some of these species, it is doubtful that a functional
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
443Physiol Rev VOL 93 JANUARY 2013 www.prv.org
peptide is produced. Even in humans where mRNA expres-
sion is detected in multiple tissues, there is no evidence for
functional peptide production. It is likely that RLN1 has
arisen as a result of a gene duplication event.
The RLN2 gene on the other hand (246) produces human
relaxin-2 in the corpus luteum of the ovary and is released to
circulate in the blood during pregnancy (47, 470), although it
is also produced by the prostate in males (259). RLN2 mRNA
is expressed in the corpus luteum, endometrium, decidua, pla-
centa, and mammary gland as well as in the heart and brain
(42). The RLN2 gene is equivalent to the RLN1 gene found in
nonprimate species, and it encodes the relaxin peptide that
was first discovered due to its roles in reproduction. The lo-
cations of sites of relaxin expression (see sect. V, FIGURE 4)
provide a clue to the widespread roles of the relaxin-RXFP1
system extending well beyond its original roles in reproduc-
tion. This section will outline the physiological roles of
relaxin acting through its receptor RXFP1 focusing pre-
dominantly on studies in rodents and correlating this with
proposed physiological roles in humans. It should be noted
that the physiological roles of relaxin outlined in this sec-
tion might not be relevant to other species. A detailed out-
line of the physiological roles of relaxin in all mammalian
species can be found here (47).
1. Reproductive physiology in the female
Relaxin produced by the corpus luteum and/or placenta has
important roles in pregnancy and parturition and is a major
circulating hormone during pregnancy in all mammalian
species. Relaxin has actions on the pubic symphysis, cervix,
uterus, vagina, and mammary glands. It also has important
roles in the cardiovascular changes that occur during preg-
nancy.
Modification to the pelvic girdle involving growth of the
interpubic ligament is essential for successful birth in many
species (47, 470, 471). Relaxin is believed to mediate the
increased flexibility and elasticity of the interpubic ligament
that is associated with pregnancy in several species (394,
496). In the relaxin knockout mouse, the interpubic liga-
GFP-βarr1 GFP-βarr2
0 min 60 min 0 min 60 min
RXFP1
RXFP2
AT1R
FIGURE 11. RXFP1 and RXFP2 receptors do not internalize following prolonged exposure to relaxin. HEK-
293 cells transiently expressing RXFP1 (Top panels), RXFP2 (Middle panels), AT1R (Bottom panels), and
green fluorescent protein (GFP)-labeled
-arrestin-1 (GFP-
arr1; left) or GFP-
-arrestin-2 (GFP-
arr2; right).
Cells were stimulated with 1
M agonist [relaxin (A); INSL3 (B); ANG II (C)] and imaged before stimulation and
60 min after stimulation. Bar indicates 5
m. Many G protein-coupled receptors such as the AT1 receptor
interact with
-arrestins and are internalized (bottom panels). RXFP1 (top panels) and RXFP2 (middle panels)
did not interact with
-arrestins 1 or 2 or internalize. [Modified from Callander et al. (87).]
BATHGATE ET AL.
444 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
ment fails to develop, suggesting a role for relaxin (588).
The mechanisms likely involve actions on collagen remod-
eling as in other reproductive organs (see below). Although
such a role is not clear in humans or other primates (339), it
has been postulated that relaxin levels correlate with severe
pelvic pain and excessive joint laxity seen in some women
during pregnancy (335, 340).
Relaxin is also responsible for the softening and hypertro-
phy of the cervix during the second half of pregnancy in
mammals, which involves effects on collagen, elastin, pro-
teoglycans, and glycosaminoglycans (47). Cervical develop-
ment is impaired in both relaxin-deficient rats (83) and
relaxin knockout mice (588), highlighting a role of relaxin
in this process. The mechanisms involve interactions with
steroid hormones and prostaglandins (47). Studies in rats
demonstrate that relaxin promotes growth of epithelial and
stromal cells in the cervix (82, 83) by stimulating cell pro-
liferation (305) and inhibiting apoptosis (590). In late preg-
nant relaxin-deficient rats, pigs, and relaxin knockout mice,
there is no dispersion or disorganization of collagen fiber
bundles in the cervix as occurs in control animals (304, 327,
589). In humans, relaxin levels increase during cervical rip-
ening, but ripening still occurs following embryo transfer
where circulating relaxin levels are undetectable (139). Al-
though direct application of porcine relaxin to the cervix
appeared in early studies to assist in ripening (152, 339),
clinical trials with recombinant human relaxin-2 failed to
confirm this finding (52).
Relaxin influences uterine contractility (295) and uterine
growth (495) during pregnancy, but this role is highly species
dependent. In humans, relaxin has little effect on uterine tone
(337, 338), but in rat, mouse, guinea pig, hamster, and pig, the
uterus is relaxed by relaxin (for review, see Ref. 47). In rat and
mouse, the uterus contains relaxin binding sites and mRNA
(47) in the myometrium (see sects. IVA1 and VA1).
In nonpregnant rats, the effect of relaxin on the growth and
development of the uterus is associated with vascular dila-
tation (547) and in rhesus and marmoset monkeys with
endometrial angiogenesis (145, 185). In primates and rats,
pretreatment with estrogens enhances the effects of relaxin
(1, 145), and in pigs, estrogens are obligatory for the
growth-promoting effects of relaxin (242, 581). The mech-
anism involved in the interaction between relaxin and es-
trogen is not well understood, although in rats estrogen
pretreatment dramatically increases relaxin receptor bind-
ing sites in the uterus (400, 401, 515). Although these find-
ings largely support a uterotrophic effect of relaxin, this
0.02
0.01
0.00
-0.01
0.02
0.01
0.00
-0.01
6810
4
2
-2 0
-4
6810
4
2
-2 0
-4
time (min)
time (min)
Ligand-induced BRET ratioLigand-induced BRET ratio
relaxin
relaxin-3
Ligand or vehicle
addition
Ligand or vehicle
addition
Interactions with β-arrestin 1
Interactions with β-arrestin 2
Unstimulated Vehicle
relaxin-3 relaxin
relaxin
relaxin-3
FIGURE 12. RXFP3 receptors interact with
-arrestins and undergo internalization. The panels show
HEK293 cells transfected with GFP2-labeled RXFP3 receptors. Unstimulated cells or cells incubated with
vehicle show RXFP3 receptors localized to the plasma membrane (top panels). Exposure of cells to human
relaxin-3 but not human relaxin-2 caused internalization of RXFP3 receptors (bottom panels) (545). Examina-
tion of interactions between RXFP3-Rluc8 and one of the
-arrestin fusion proteins (
-arrestin 1-Venus or
-arrestin 2-Venus) cotransfected into Flp-In CHO cells using BRET showed that interactions occurred between
RXFP3 and both
-arrestin 1 and
-arrestin 2 following human relaxin-3 but not human relaxin-2. Scale bar
10
m.
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
445Physiol Rev VOL 93 JANUARY 2013 www.prv.org
does not play a role in pregnancy in mice and rats (252) but
is important in pigs (365). In humans it is likely that utero-
trophic effects of relaxin are associated only with implan-
tation (see sect. VIIIA2) and not with the later stages of
pregnancy.
Relaxin also influences the growth of the vagina during
pregnancy in mice (194, 450) and rats (82, 590, 591). In
relaxin knockout mice, this growth does not occur, high-
lighting the essential role of relaxin (589). Relaxin binding
is present in the vaginal luminal epithelial cells as well as
circular and longitudinal smooth muscle cells in the rat
(591) and human (288). The effect of relaxin on vaginal
development during pregnancy together with its actions on
the cervix and pubic symphysis are essential for normal
delivery in some species.
Relaxin has trophic effects on the mammary gland. In
mice transgenically overexpressing mouse relaxin, there
is a 20-fold increase in serum levels that is associated
with hypertrophic nipple development in virgin females
(158). This phenotype is absent in mice that also had
deletion of RXFP1 (158). In rats (205) relaxin in combi-
nation with estrogen and progestagen promotes mam-
mary gland development in ovarectomized immature an-
imals. The major effect in rodents is on nipple develop-
ment, whereas in pigs it is mainly on the mammary gland
(248). In rats, administration of a monoclonal antibody
to relaxin during the second half of pregnancy prevents
nipple development (251), and an identical phenotype
occurs in relaxin knockout mice (588). Pups born to
relaxin knockout mice die within 24 h unless cross-fos-
tered to wild-type mothers. The effects are entirely due to
the poor nipple development as female knockout mice are
able to produce milk normally (588). The same pheno-
type is displayed by RXFP1 knockout mice (293) and is
not rescued by administration of INSL3 (274). In con-
trast, relaxin-deficient pigs show normal nipple develop-
ment and function (248, 582). Mammary gland develop-
ment in mice and rats does not require relaxin, although
it may affect tissue differentiation, but in pigs the devel-
opment of the mammary gland parenchyma requires re-
laxin. Relaxin binding sites indicative of RXFP1 recep-
tors are present in the mammary glands of pigs, rats, and
humans. In humans, RXFP1 receptors are localized to the
nipple, epithelial cells (288), and stromal tissue (255).
Major cardiovascular changes are associated with preg-
nancy in most mammals. Relaxin directly acts on the
blood vessels, kidney, and heart to contribute to the car-
diovascular changes that occur during pregnancy (see
sect. VIIA4). In rodents, relaxin causes a decrease in
plasma osmolality during pregnancy associated with
changes in thresholds for thirst and AVP secretion (311,
387). In relaxin-deficient rats (387) and relaxin knockout
mice (588), the decrease in plasma osmolality does not
occur. Although plasma osmolality is reduced in human
pregnancy, there is currently no evidence that relaxin is
involved. Relaxin may also influence water drinking dur-
ing pregnancy in rodents.
Relaxin has many well-defined reproductive roles in many
species, but in humans these effects are often absent or
ill-defined. Relaxin is produced in the human ovary during
the follicular phase of the menstrual cycle and is secreted
into the circulation (99, 140, 497). It has been suggested
that this relaxin facilitates implantation and maintenance of
pregnancy in the first trimester. Early studies showed that
administration of relaxin to monkeys caused growth of, and
increased angiogenesis in, the uterus (145, 185, 214, 222).
It is likely that similar effects occur in the human endome-
trium, since in the clinical trial of relaxin for the treatment
of scleroderma (149), women who received 24-wk subcu-
taneous infusion of relaxin reported heavy, irregular, or
prolonged menstrual bleeding (540). However, in humans
and other primates, the peak of relaxin secretion in the first
trimester of pregnancy coincides with embryo implanta-
tion, suggesting that relaxin is involved in implantation (see
sect. VIIIA2). Although there is an association between re-
laxin secretion and the success of implantation in primates,
it is clearly not mandatory since women without ovaries can
become pregnant by ovum donation even though they have
undetectable levels of circulating relaxin (267). Relaxin secre-
tion is correlated with increased expression of RXFP1 mRNA
and relaxin binding in the human endometrium in the secre-
tory phase of the menstrual cycle (67, 88). RXFP1 immunore-
activity is also shown in endometrial stromal cells (255) in
humans and monkeys (145). Relaxin also induces decidualiza-
tion of endometrial stromal cells in culture (529). In contrast,
relaxin does not appear to be important for the decidualiza-
tion of the endometrium in mice (168).
In humans and other primates, increasing evidence points to a
role for relaxin in implantation. Relaxin treatment is associ-
ated with increased endometrial angiogenesis, thickening, and
bleeding (145, 185). In macaques, relaxin treatment of cycling
females in the peri-implantation period increases endometrial
thickness and implantation-related bleeding. Relaxin treat-
ment also increased rates of implantation and multiple preg-
nancies (214). In addition, plasma levels of relaxin peak in the
first trimester, coincident with embryo implantation. Rises in
relaxin levels are associated with chorionic gonadotrophin
(497, 498), are altered in patients with early pregnancy loss
(498), and are predictive of in vitro fertilization success in
granulosa cell cultures (499). However, relaxin probably only
facilitates implantation since this still occurs in humans and
other primates that lack ovaries (182, 267). The relaxin that is
produced in the male reproductive tract appears in semen and
increases sperm motility and facilitates penetration into
oocytes (557). A potential role for this relaxin may be to act on
the female reproductive tract to prepare the endometrium for
implantation.
BATHGATE ET AL.
446 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
2. Reproductive physiology in the male
Relaxin is found in the male reproductive tract in most
mammals. Human relaxin-2 has been purified from human
male seminal plasma and is identical to luteal relaxin (567).
The source of this relaxin appears to be the prostate (489,
576), but its function in the male is unknown, although as
discussed above seminal relaxin may be involved in implan-
tation. RXFP1 has been localized to sperm in both mice
(293) and humans (91, 160), and treatment of human
sperm with relaxin increases motility (91) and induces hy-
peractivation, intracellular calcium and cAMP, and acro-
some reaction in human sperm (160).
Prostate-derived relaxin may have local actions in the male
reproductive tract, since the testis and prostate in mice (442),
rats (237), and humans (239) express RXFP1 mRNA. Inter-
estingly, studies in one strain of the relaxin and RXFP1 knock-
out mouse suggested a role for relaxin in male reproduction.
The mice displayed poor growth of the reproductive tract with
the relaxin knockout, demonstrating effects on the epididymis,
seminal vesicles, and prostate (442), whereas the RXFP1
knockout mice displayed a smaller epididymis (293). Reduced
growth was associated with increased collagen as demon-
strated in the reproductive organs of the female relaxin knock-
out mouse (589). Examination of sperm maturation in the
relaxin and RXFP1 knockout animals suggested that there
might be apoptosis in early stages of spermatogenesis (293,
442). In this relaxin knockout mouse strain (442) and one
strain of RXFP1 knockout mice (293), the changes in the male
reproductive tract are associated with reduced fertility. How-
ever, independent strains of Rln1 /and RXFP1 knockout
mice (180) showed no prostate phenotype, and another strain
of RXFP1 knockout mice had no male reproductive tract phe-
notype or association with reduced fertility (274). There are
also reports that the phenotypes seen in early generations of
relaxin and RXFP1 knockout are lost in later generations
(257), a similar phenomenon to that seen in the INSL6 knock-
out (85). The authors suggest that the phenotype may “be
taken as evidence of a genuine physiology, but one which
because of redundancy, genetic modifiers, and possibly trans-
generational epigenetic modification can be masked by in-
breeding” (257). However, the contradictory results from
other strains of relaxin and RXFP1 knockout mice suggest
that a clear role for relaxin in the male reproductive tract
awaits further studies.
3. Central actions of relaxin on RXFP1
In mammals, relaxin acts directly on receptors located in the
SFO and OVLT to cause a reduction in plasma osmolality
(504). In rats, the decline in plasma osmolality during the
second half of pregnancy is associated with increased serum
relaxin levels (311, 472) and is absent in pregnant relaxin-
deficient rats that have undergone ovariectomy or been
treated with relaxin antibodies (387). Likewise, in wild-
type mice, the decrease in plasma osmolality that occurs in
late pregnancy is not observed in relaxin knockout mice
(588). In humans, there is a decrease in plasma osmolality
with pregnancy, but this may not be an effect of relaxin,
since women that become pregnant following ovum dona-
tion have undetectable relaxin levels yet still show this effect
(267, 268). However, the apparent lack of relaxin in these
conditions may have to be reassessed using more sensitive
methods of detection since the discovery of responses to
relaxin (201) at plasma levels well below the level of detec-
tion using current ELISA techniques. In rats, water con-
sumption is strongly stimulated by relaxin and is also in-
creased in the second half of pregnancy (396, 502). Intracere-
broventricular (ICV) or intravenous relaxin also promotes
drinking in nonpregnant rats (502). In rats given relaxin
monoclonal antibodies in the second half of pregnancy, there
is a reduction in water consumption. These actions are pro-
duced by relaxin acting on RXFP1 receptors located in the
SFO and OVLT (FIGURE 6). Administration of porcine or hu-
man relaxin-2 (intravenously) causes increased c-fos expres-
sion in neurons of the peripheral and dorsal segments of the
SFO and in the dorsal cap region of the OVLT, as well as in the
supraoptic and paraventricular nuclei of the hypothalamus
(361, 362, 504, 505), all sites of localization of RXFP1 (FIG-
URE 6). Circulating relaxin penetrates these brain regions and
may be responsible for the plasma osmolality changes that
occur during pregnancy (556).
Other central actions of relaxin have been studied in
much less detail. RXFP1 receptors in the circumventricu-
lar organs and hypothalamic nuclei may have a role in the
timing of parturition in rats since this is disrupted by
central administration of a relaxin monoclonal antibody
(503). RXFP1 receptors are highly expressed in the ba-
solateral amygdala and administration of relaxin to this
region impairs fear related memory consolidation in rats
(331). However, no specific agonist or antagonist studies
have been carried out to determine whether these effects
are mediated by RXFP1 or to determine the source of
endogenous relaxin that activates these receptors (86).
Although RXFP1 is also highly expressed in other regions
associated with memory formation such as the neocor-
tex, thalamic nuclei, hippocampus, and supramammil-
lary nucleus (FIGURE 6), there are no studies to date that
have examined possible effects on memory. RXFP1 re-
ceptors are also present in the oxytocin-containing cells
of the paraventricular and supraoptic hypothalamic nu-
clei (81), and relaxin administration increases oxytocin
neuron activity and oxytocin release (555).
4. Effects of relaxin on the cardiovascular system
Relaxin has an important role in many of the adaptive car-
diovascular changes that occur in pregnancy (101). These
include increases in plasma volume, cardiac output, and
heart rate, as well as decreased blood pressure and vascular
resistance (103, 104, 118, 119). In both female and male
rats, relaxin administration increases renal plasma flow and
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
447Physiol Rev VOL 93 JANUARY 2013 www.prv.org
glomerular filtration rate (114) and in female rats causes the
reduction in plasma osmolality associated with pregnancy
(387). In rats, ovariectomy or passive immunization with
monoclonal antibodies for rat relaxin prevent these adap-
tive changes (387). In humans, relatively few studies have
been conducted; however, they do suggest similar effects of
relaxin on the cardiovascular system. In the clinical trial for
scleroderma, long-term (6 mo) infusion of relaxin increased
creatinine clearance and produced a modest decrease in
blood pressure (149, 527). More recent trials of relaxin for
the treatment of cardiac failure have shown that a short
(24 h) infusion of relaxin is associated with decreased sys-
temic vascular resistance, serum creatinine, pulmonary
wedge pressure, and a small decrease in systolic blood pres-
sure (133, 134, 527). Comparison of the cardiovascular
changes that occur in pregnancy with those seen following
administration of relaxin shows marked similarity in both
rats and humans (105).
Vasodilation in arterioles, capillaries, and venules is a com-
mon response to relaxin in reproductive tissues (35, 304,
547), heart (33, 36, 343), liver (34), and cecum (58). Re-
laxin is a potent vasodilator in arteries (102, 360), although
the effect is vessel specific (360). Relaxin is a physiological
antagonist of vasoconstrictors in mesenteric arteries (348,
492), primary bovine aortic smooth muscle cells (31), and
uterine artery (318). Relaxin produces its vasodilator ef-
fects in guinea pig and rat coronary arteries by increasing
NO synthesis (36), later confirmed in bovine cultured
smooth muscle cells (31). The hormone also reduces the rise
in intracellular Ca
2
produced by
-thrombin or angioten-
sin II (31, 154). In humans, relatively few studies have been
conducted, but those performed to date show vasodilator
effects in gluteal resistance or subcutaneous arteries but
little or no effect in pulmonary, myometrial, or placental
vessels (170, 360, 415). In gluteal arteries, the vasodilator
responses likely involve NO and interestingly were influ-
enced by the medication being administered to patients.
Arteries obtained from patients on ACE inhibitors showed
marked attenuation of the vasodilator response to relaxin,
effects that appeared to be further enhanced by inhibition of
cyclooxygenase (170). However, in patients not receiving
ACE inhibitors, indomethacin had little effect (170). The
vasodilator mechanisms suggested for relaxin in both hu-
mans and in animal models involve activation of NOS (104,
385), VEGF, matrix metalloproteinases, ET
B
receptors
(129, 389), and modification of the extracellular matrix of
the vessel walls (FIGURE 9) (263, 307, 360, 568).
The molecular mechanisms of relaxin-induced vasodilation
depend on the duration of hormone exposure, that is, there
are rapid and sustained vasodilatory responses. Our current
understanding is that the vasodilatory responses to relaxin
are mediated by actions at its major receptor RXFP1. Rapid
relaxation responses to relaxin have been recorded in hu-
man gluteal arteries (170) and rat and mouse renal but not
rat mesenteric arteries (360). The responses are endothe-
lium dependent and blocked by NOS inhibitors, the PI3K
inhibitors wortmannin and LY294002 and by PTX pre-
treatment but not by the VEGF receptor antagonist SU5416
(FIGURE 9) (360). The current data suggest that the re-
sponses involve
␤␥
activation of PI3K, Akt phosphoryla-
tion, and eNOS. Sustained responses to relaxin also involve
NOS, since blockade with L-NMMA prevents the renal he-
modynamic and hyperfiltration responses to relaxin (114),
and these effects can be recapitulated in vitro (389). Treat-
ment with NOS inhibitors or removal of the endothelium
also enhances contractions of vascular smooth muscle to
agonists in both pregnant rats (111, 179) and in rats treated
chronically with relaxin (389). The mechanism involved
does not appear to be attributable to changes in NOS ex-
pression, since the alterations in renal haemodynamics as-
sociated with pregnancy are not accompanied by large
changes in NOS expression. In pregnant rats, eNOS expres-
sion in the renal artery fell by 39%, whereas iNOS and
nNOS expression increased by 31 and 25%, respectively
(5). Rather, it is likely that the vasodilator response involves
endothelin acting at endothelial ET
B
receptors to release
NO (FIGURE 9). This could be further enhanced as a result
of increased expression of the ET
B
receptor (129), although
other studies have failed to find support for this mechanism
(279). What is clear, however, is that the ET
B
receptor
antagonist RES-701–1 blocks the renal hemodynamic
changes produced by relaxin (113) in rats and antagonizes
inhibition of renal artery smooth muscle produced by re-
laxin or in pregnancy (389). Similar effects are produced in
vitro by the ET receptor antagonist SB209670 but not by
the ET
A
selective BQ123 (179).
There is substantial evidence to suggest that the ligand that
interacts with ET
B
receptors to activate NOS is ET
1–32
,
produced by the action of MMP2 and MMP9 on big ET
(163). The link to relaxin is that in blood vessels from
pregnant or relaxin-treated nonpregnant rats, pro-MMP2
and MMP2 activity and pro-MMP2 protein and mRNA are
increased (262, 263). MMP9 activity also appears to be
increased somewhat, although more recent studies suggest
that MMP9 is more important in relatively short-term re-
sponses to relaxin (46 h) with reversal of the effects being
produced by MMP9 rather than MMP2 neutralizing anti-
body (264). The involvement of MMP2 and MMP9 in the
vasodilator responses to relaxin or pregnancy is supported
by studies with MMP inhibitors. The selective MMP2 in-
hibitor cyclic CTTHWGFTLC, the MMP inhibitor
GM6001, TIMP-2, and MMP-2 neutralizing antibody all
inhibit the vasodilator actions of relaxin in renal arteries
(263), whereas inhibition of the formation of ET
1–21
by
phosphoramidon had no effect. These studies clearly impli-
cate MMPs in the vasodilator actions of relaxin, although it
should be borne in mind that MMPs also influence vasore-
activity by interacting with other systems such as cleavage
BATHGATE ET AL.
448 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
of calcitonin gene-related peptide to promote vasoconstric-
tion (164).
While a case has been made for a role of VEGF in the
vasodilator actions of relaxin, it is not yet precisely clear
what that role is. In nonvascular tissues, treatment with a
range of compounds that increase cAMP levels also in-
creases VEGF expression. So, in human endometrial cells,
relaxin increases VEGF expression, and these effects are
prevented by AC inhibition, and mimicked by either the AC
activator forskolin or a PDE inhibitor (540), suggesting that
increased cAMP production mediates increased VEGF tran-
scription, and thus angiogenesis (FIGURE 9). However, in
vascular tissues, the effects are less clear. Preincubation of
rat or mouse renal arteries or human subcutaneous arteries
with the VEGF receptor antagonist SU5416 blocks the va-
sorelaxant effects of relaxin (359), although the effects of
SU5416 alone and its specificity can be questioned (17,
316). In other studies using rat renal arteries, SU5416 po-
tentiated the vasodilator effects of relaxin (360). VEGF neu-
tralizing antibodies also block the effects of relaxin, but this
can be mimicked using placental growth factor antibodies
(359), again making interpretation of the results difficult. It
is possible that the antibodies by blocking the vasodilator
effects of VEGF and PGF may produce physiological antag-
onism of the response to relaxin without necessarily being
directly related to the primary mechanism of action.
Relaxin also acts directly on the heart. The presence of
relaxin receptors (RXFP1) in the heart was first suggested
by the demonstration of high-affinity binding sites for re-
laxin in rat atria (401, 402). Subsequently, in the same
species, relaxin was shown to be one of the most powerful
inotropic and chronotropic agents known (271). The posi-
tive chronotropic effects of relaxin occur in perfused intact
hearts (36, 107, 531, 534) and isolated right atria (271,
350, 514, 553, 554), and the positive inotropic effects occur
in left atria (271, 350, 514, 553, 554). The chronotropic
effects of relaxin occur together with the secretion of atrial
natriuretic peptide in isolated perfused rat hearts (534). In
rat atrial myocytes, relaxin inhibits outward potassium cur-
rents, increases action potential duration, and enhances
Ca
2
entry (417, 418). In rabbit sinoatrial node cells, re-
laxin caused increases in the rate of spontaneous action
potentials and increased the L-type Ca
2
current (206) by a
PKA-dependent mechanism. Until recently, these actions of
relaxin on the heart were thought to be largely confined to
rodents. However, a recent study (127) shows that relaxin
has similar inotropic effects in human atria that are pre-
served in failing hearts and involves PKA, outward K
cur-
rents, and PI3K. Interestingly, the effects of relaxin in the
cardiovascular system, unlike the effects on reproduction,
are not gender specific and are observed in both males and
females. There is also evidence that relaxin protects against
myocardial injury caused by ischemia and reperfusion (33).
Pretreatment of rats with relaxin 30 min before 30 min of
cardiac ischemia produced by ligating the left anterior de-
scending coronary artery markedly reduced the size of the
penumbra, and reduced cardiac arrhythmias, mortality as
well as myeloperoxidase activity, malonyldialdehyde pro-
duction, Ca
2
content as well as causing an improved mor-
phology (33). Relaxin may be a naturally occurring cardio-
protective agent, since in congestive heart failure the expres-
sion of human relaxin-1 and human relaxin-2 is increased in
both atria and ventricles and the level of expression follows
the degree of failure (132, 169) but is not a predictor of
clinical outcomes (169).
5. Effects of relaxin-RXFP1 on connective tissue
metabolism
The effects of relaxin on collagen synthesis and breakdown
in reproductive tissues were the first biological effects of
relaxin to be recorded (220). Since then it has become evi-
dent that relaxin has more general antifibrotic properties
(47, 181), and a number of attempts have been made to put
these to therapeutic use (462, 526–528, 538). One of the
interesting aspects that has emerged is that the antifibrotic
properties of relaxin are clearly seen only in disease condi-
tions associated with excessive collagen deposition. Several
studies have examined relaxin as a possible treatment for
the connective tissue disease scleroderma. Although relaxin
was shown to be safe and well-tolerated in clinical trials,
and even effective in some patients in a phase II trial (462),
it failed to show clinical efficacy in a larger scale phase III
trial (149). In spite of these disappointing findings, there has
been increasing evidence from animal studies that relaxin
has a role in controlling collagen turnover. In relaxin
knockout mice, there is a progressive increase in tissue fi-
brosis with age in male mice that is reversed in lung (444),
kidney (445), and heart (443) by the administration of re-
laxin.
In the lung, TGF-
causes detrimental changes associated
with fibrosis that are reversed by relaxin. Treatment with
relaxin reduces expression of collagen types I and III and
increases levels of MMPs (542). Antifibrotic effects were
also produced in mice treated with bleomycin (542). In
kidney-derived fibroblasts, relaxin displays similar effects
on profibrotic changes induced by TGF-
and was shown to
mediate these effects through the NO/guanylyl cyclase
pathway that caused decreases in Smad2 phosphorylation
and nuclear localization (FIGURE 10) (370). Relaxin also
displays antifibrotic effects in a variety of rat renal models
including fibrosis produced by bromoethylamine treatment
(181), in an anti-glomerular basement membrane model
(356), and in spontaneously hypertensive rats (307). In car-
diac fibroblasts, relaxin reduces collagen type I and III ex-
pression and increases MMPs (443). In cardiac fibrosis pro-
duced by chronic stimulation of
-adrenoceptors by isopren-
aline (585) or by cardiac specific transgenic overexpression of
2
-adrenoceptors (44), relaxin administration or delivery
via an adenovirus vector markedly reduced cardiac fibrosis.
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
449Physiol Rev VOL 93 JANUARY 2013 www.prv.org
In streptozotocin-treated mRen-2 rats (a model for diabetic
cardiomyopathy), relaxin treatment reduced left ventricular
collagen, myocardial stiffness, and diastolic dysfunction
(440). This was associated with a significant decrease in
TIMP-1 expression and an increase in extracellular matrix-
degrading MMP-13 (440). However, in a chronic pressure
overload model in mice, relaxin was ineffective (569), pos-
sibly because cardiac RXFP1 may be downregulated in this
model or the serum concentrations of relaxin are insuffi-
cient to offset the extensive fibrosis. There is also good
evidence supporting antifibrotic actions of relaxin in a num-
ber of models of liver fibrosis (54). Overall, the evidence
suggests that relaxin may have useful antifibotic properties
in a number of pathological conditions.
In a recent study it has been shown that relaxin treatment of
primary fibrochondrocytes increases the expression of
mRNA for MMP-9 and MMP-13 and that the effect is
mediated by RXFP1 and not RXFP2 receptors (4). Similar
treatment caused activation of PI3K, Akt, PKC-
, and
ERK1/2 that could be blocked by the use of appropriate
inhibitors that also blocked the induction of MMP-9. The
effect of relaxin on MMP-9 expression was also blocked by
prior transfection of a dominant negative form of Akt or by
siRNA knockdown of ERK1/2, PKC-
, Elk-1, c-fos, and to
a minor extent NF
B (4). This is one of the first studies that
connects many of the known pathways of relaxin/RXFP1
signaling to a recognized response to relaxin treatment and
provides insights that could be useful in translating the an-
tifibrotic effects of relaxin into a clinical setting.
Relaxin also has beneficial effects in wound healing (92). This
may involve the vasodilator effects mentioned above, but in
addition, relaxin may hasten the synthesis of new blood vessels
by enhancing the local production of VEGF (541).
6. Role of relaxin in the formation and spread of
tumors
There is also now increasing evidence that relaxin is pro-
duced by cancer cells and can act in an autocrine manner
on RXFP1 receptors expressed on these cells. To date,
relaxin has been shown to be expressed by endometrial
(273), mammary (522), thyroid (226), and prostate tu-
mors (157, 532). There has long been an association of
relaxin with breast cancer (reviewed in Refs. 26, 479),
and relaxin treatment of breast cancer cells increases
their invasive potential (59). Furthermore, elevated se-
rum relaxin levels have been reported in breast cancer
patients and in patients with metastases (60). It is possi-
ble that the relaxin produced by breast cancer cells is
involved in tissue remodeling during breast cancer pro-
gression (60). Relaxin has also been associated with pros-
tate cancer progression, and blocking the actions of re-
laxin or RXFP1 in rodent models of prostate cancer re-
sults in decreased cancer growth (see sect. VIIIA7).
B. INSL3-RXFP2
INSL3 was isolated from a porcine testis cDNA library and
designated Leydig insulin-like peptide (2). Murine INSL3
was independently cloned from a cDNA library of clones
preferentially expressed in testicular tissue (424). INSL3 is a
major secretory product of the prenatal and postnatal tes-
ticular Leydig cells in all species tested (2, 424). Detection of
INSL3 expression in human cyclic corpus lutea put to rest
suggestions that INSL3 expression is restricted to testis
(521). Transcripts were subsequently detected in other tis-
sues and species including ruminant ovary, uterus, and pla-
centa (40, 432) as well as mouse (595) and marmoset ovary
(583).
1. Reproductive physiology: male
The development of INSL3 knockout mice enabled physio-
logical functions to be attributed to the peptide (379, 596).
INSL3 knockout mice develop normally, but male knock-
out mice are infertile and bilaterally cryptorchid, with the
testis located high in the abdominal cavity adjacent to the
kidney. There may also be torsion of the vas deferens and
testicular artery and localization of the right testis in the
contralateral position (379) likely due to a lack of attach-
ment of the testis to the inguinal region by the gubernacu-
lum and regression of the suspensory ligament. At birth, the
size and histology of testes from knockout animals was
normal but subsequently degenerated, and spermatids,
spermatozoa, and mature sperm were absent in adults. This
is likely to be a temperature effect, since surgical reversal of
cryptorchidism led to normal spermatogenesis in seminifer-
ous tubules (384, 596).
The cryptorchid phenotype is not due to androgen defi-
ciency, and INSL3 knockout males have serum testosterone
levels similar to wild-type and heterozygous males (379)
and apart from cryptorchidism have normal genitalia (596)
and androgen-dependent behaviors (379). The gubernacu-
lum in knockout mice is similar to that found in wild-type
females. In vitro, coculture of explant gubernaculum from
either wild-type or knockout mice, with the testis, induced
growth of the gubernaculum (147), suggesting that INSL3
secreted by the testes is important for testicular descent.
Similarly cryptorchidism is the phenotype displayed in the
white spotting (crsp) transgenic mouse which has a 550-bp
deletion affecting the G protein-coupled receptor affecting
testis descent (GREAT) (186, 406). This receptor (formerly
LGR8) is now termed RXFP2, the cognate receptor for
INSL3 (42). Since gubernacular cells undergo cell division
in response to INSL3, both in vitro and in vivo studies
suggest that the INSL3-RXFP2 system in male urogenital
structures is essential for development of the gubernaculum
during embryogenesis and for normal transabdominal tes-
ticular descent. A recent study has demonstrated that
INSL3 signaling in the fetal gubernaculum both in vitro and
BATHGATE ET AL.
450 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
in vivo involves genes that induce morphogenetic programs,
including BMP and WNT signaling (266). Additionally,
another study demonstrated that INSL3 signaling is essen-
tial for myogenic differentiation in the gubernaculum and
involved genes in the Wnt/
-catenin and NOTCH path-
ways (270).
Transgenic mice overexpressing INSL3 were generated to
further investigate the actions of INSL3 on the gubernacu-
lum (3, 292). Females display ovaries positioned over the
bladder, attached to the abdominal wall by well-developed
craniosuspensory ligament (CSL) and gubernacula. Al-
though this supported the idea that gubernacular develop-
ment is an androgen-independent process, subsequent stud-
ies suggested otherwise (see below). In transgenic males, the
pancreatic expression of INSL3 rescued the cryptorchid
phenotype resulting from the deletion of the endogenous
testicular INSL3 gene. Finally, in female INSL3 transgenic
RXFP2 knockout mice, there is a wild-type phenotype, con-
firming that the actions of INSL3 on the gubernaculum are
mediated by RXFP2 (65).
However, there is a suggestion that INSL3 alone is not
sufficient for normal testicular descent. Normal testicular
descent appears to rely on both INSL3-induced develop-
ment of the gubernaculum and androgen-mediated cranial
suspensory ligament regression (3, 375). Androgens have
been shown to modulate RXFP2 expression in the guber-
naculum. Androgens alone are insufficient to stimulate gu-
bernacular cell proliferation but in combination with
INSL3 have a proliferative effect, which is blocked by an
RXFP2 antagonist or siRNA (580). In vivo,the RXFP2
antagonist blocked testosterone replacement therapy-in-
duced testis descent in 50% of LH receptor knockout mice.
Thus RXFP2 signaling appears to mediate the effects of
androgens on the gubernaculum, which is important for the
inguinoscrotal phase of testicular descent. This is particu-
larly important as most human defects in testicular descent
are associated with the inguinoscrotal phase (250).
Many early papers investigating the role of INSL3-RXFP2
in cryptorchidism emphasized the relevance of this ligand-
receptor pair to cryptorchidism in humans. Cryptorchidism
is the most common birth defect of the male genitalia, af-
fecting 1–4% of live male births, with a greater incidence in
premature infants (257). While the phenotype of INSL3 or
RXFP2 knockout mice suggested that mutations in these
genes could account for cryptorchidism in some infants, this
is in fact highly unlikely. Some cryptorchid patients display
polymorphisms in either INSL3 or RXFP2, but very few of
these mutations confer a functional change in either the
peptide or the receptor. Mutation detection analysis in
genomic DNA samples from formerly cryptorchid patients
identified two mutations in INSL3 (533). The first a non-
sense mutation in the C-peptide region of pro-INSL3
(R49X) results in a nonfunctional peptide, and the second,
in the B chain (P69L) of the pro-INSL3 sequence results in a
peptide with reduced activity (65). Similarly, of the four
RXFP2 mutations identified in humans, only one is ex-
pected to alter receptor function. The T222P mutation in
the fourth leucine-rich repeat is thought to result in a con-
formational change that interferes with ligand action (186),
probably by inhibiting expression of mutant RXFP2 recep-
tors at the cell surface (64). Recently, an N-ethyl-N-nitro-
sourea-induced mutation in mice that impaired testicular
descent and reduced testis weight was found to be due to a
mutation in leucine-rich repeat 8 (D294G) (209). Recom-
binant expression of the RXFP2 D294G mutant showed
reduced cell surface expression and a corresponding reduc-
tion in maximal cAMP response compared with wild type.
Patients with the T222P mutation displayed heterogeneous
phenotypes such as unilateral and bilateral cryptorchidism,
retractile testis, normozoospermia, and complete azoosper-
mia (186). All mutations of INSL3 and RXFP2 identified in
cryptorchid patients have only been found in the heterozy-
gous state. As the expression of INSL3 in testicular Leydig
cells is exceptionally high, it is unlikely that a mutation in a
single allele would be sufficient to induce bilateral cryp-
torchidism.
Nonetheless, there is evidence that the INSL3-RXFP2 sys-
tem is affected in cryptorchid patients that have been ex-
posed to environmental toxins. The use of diethylstilbestrol
for hormonal support in pregnant women was terminated
due to a high incidence of genital defects such as cryp-
torchidism in offspring (183). In mice, maternal exposure to
estrogens, including 17
- and
-estradiol and diethylstil-
bestrol, downregulates expression of INSL3 in embryonic
Leydig cells, and the gubernacula fail to develop which
results in cryptorchidism (380). Hence, it is possible that
environmentally derived estrogens could cause cryptorchid-
ism in humans by suppression of INSL3 expression in the
fetal testis. Similarly, treatment of pregnant rats with phtha-
late esters reduces testicular INSL3 expression as a conse-
quence of Leydig cell toxicity (97). As phthalates are used in
a variety of products as plasticisers, humans in developed
countries are substantially exposed, although, due to health
concerns, their use is decreasing.
2. Reproductive physiology: female
INSL3 is produced in the ovary of all mammals predomi-
nantly in the thecal cells of the follicle (256). Female INSL3
knockout mice demonstrate impaired fertility associated
with longer estrous cycle length, smaller litter sizes, accel-
erated follicular atresia and luteolysis, and premature loss
of corpora lutea (379, 490). These phenotypes suggest that
INSL3 inhibits follicular and luteal cells from entering the
apoptotic pathway. In support of this, INSL3 expression in
vivo varies with follicular development with levels greatest
at the early antral stages and declining prior to expression
of cholesterol side-chain cleavage cytochrome P-450 or 3
-
hydroxysteroid dehydrogenase expression in the theca in-
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
451Physiol Rev VOL 93 JANUARY 2013 www.prv.org
terna as follicles enlarge or enter atresia (253). Thus thecal
cell INSL3 expression may be important in follicle selection
in addition to being antiapoptotic.
In mouse ovaries, INSL3 expression coincides with the ap-
pearance of growing follicles, and expression is higher in
the follicular than the luteal phase (595). During pregnancy,
INSL3 mRNA is greatest during phases 1–3 of follicle de-
velopment and decreases to its lowest levels between day 8
and 17 of pregnancy (595). Together, these results suggest
that INSL3 expression is correlated with follicular matura-
tion.
In rats, INSL3 may mediate the actions of gonadotropins in
both female and male germ cells. In the mammalian ovary,
oocytes show prolonged arrest at the prophase of meiosis I.
The preovulatory surge in luteinizing hormone induces the
resumption of meiosis in the oocyte, reprogramming of mu-
ral granulosa cells, expression of new mRNAs and proteins,
and changes to the secretory properties of the cells sur-
rounding the oocyte. Control of these factors is associated
with the levels of cAMP in the oocyte (535). In rat ovary,
expression of RXFP2 is restricted to the oocyte, suggesting
a similar paracrine function for the INSL3-RXFP2 system.
LH also increases transcription of INSL3 in ovarian theca
cells, and binding of INSL3 to RXFP2 expressed in the
oocyte leads to activation of G
i
and decreases in cAMP
production (276). Studies in cultured preovulatory follicles
from gonadotropin-treated female rats further indicated
that INSL3 causes meiotic progression of the arrested
oocytes.
The ability of LH to regulate INSL3 levels is supported by
data from several other species and models, including hu-
mans (50), mice (159), and roe deer (229). INSL3 may also
be involved in polycystic ovary syndrome (PCOS), a com-
mon female endocrine disorder characterized by obesity,
anovulation or amenorrhea, acne, and excessive production
or enhanced effects of androgens. In normal women, INSL3
levels are positively correlated with total testosterone, free
androgen index, basal LH concentrations and hirsuitism
score, and negatively correlated with frequency of menstru-
ation (178). In patients with PCOS, circulating INSL3 levels
are associated with higher luteinizing hormone levels, in-
creased basal and ovarian hyperandrogenemia and hirsuit-
ism score, worsened menstrual pattern, and higher mean
ovarian follicle number. It is likely that in normal-weight
women with PCOS, increased LH may be responsible for
increased androgen secretion from ovarian tissues mediated
by INSL3 (178).
3. Neurophysiology
RT-PCR was used to demonstrate RXFP2 expression in
human brain (239), and in rats, in situ hybridization histo-
chemistry showed RXFP2 mRNA in numerous thalamic
nuclei (469), including the parafascicular nucleus; the dor-
solateral, ventrolateral, and posterior thalamic nuclei; and
the medial habenula (460). These sites correspond to the
distribution of RXFP2 determined by
125
I-INSL3 binding
(460).
In contrast, INSL3 mRNA has only been detected in bovine
hypothalamus by RT-PCR (40) and Northern blotting (41)
and is not present in rodent brain (469). This may reflect the
greater importance of INSL3 in ruminants (40), which do
not express relaxin (560). The ligand-receptor mismatch
may support the proposal that RXFP1 and RXFP2 evolved
as receptors for relaxin family peptides during mammalian
evolution (561) and that prior to the emergence of INSL3,
there was a now defunct ligand-receptor pairing of which
only RXFP2 remains.
It is also possible that INSL3 derived from testicular or
ovarian cells crosses the blood-brain barrier to activate
RXFP2. Circulating INSL3 levels are relatively high (9,
172) at 1% of testosterone levels. However, INSL3 infu-
sions into rat brain suggest that RXFP2 influences sensori-
motor rather than gonadal function (460). It remains to be
determined whether INSL3 crosses the blood-brain barrier
to activate RXFP2.
4. Physiological functions of INSL3-RXFP2 in other
tissues
Knowledge of other physiological functions of INSL3 is
based largely on the developmental pattern of INSL3 and
RXFP2 expression in bone, tumors, and kidney.
INSL3 levels vary throughout life and often closely mirror
testosterone, which is also produced by the Leydig cells of
the testis. During adulthood, Leydig cell function may be
altered and INSL3 levels decline in late-onset hypogonad-
ism. In 25 young adult men with cryptorchid hypogonad-
ism caused by the T222P mutation of RXFP2, 64% had
reduced bone density but normal plasma testosterone levels
(162). RXFP2 mRNA and protein are present in human
osteoblast cells, and an osteoblast cell line responded to
INSL3 with a dose-dependent increase in cAMP. Like hu-
man, mouse osteoblasts express RXFP2, but not INSL3,
and RXFP2 knockout mice are osteopenic and have func-
tional osteoblast impairment (162). This suggests that
INSL3 may have an endocrine role in bone physiology and
that RXFP2 mutations may be linked with osteoporosis in
men. In cultured human osteoblasts (414), INSL3 regulates
the expression of genes necessary for proliferation and dif-
ferentiation, matrix deposition, and osteoclastogenesis, in
addition to dose-dependent proliferation. This correlates
well with deficits in INSL3-RXFP2 signaling that are corre-
lated with reduced bone mass (161, 162).
A role for INSL3 in tumor biology has been suggested on
the basis of expression of INSL3 and a splice variant in
human hyperplastic thyroid adenoma and thyroid cancer
BATHGATE ET AL.
452 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
(228). INSL3 is also present in human prostate carcinomas
and increases motility in a human androgen-insensitive
prostate carcinoma cell line PC-3 (286). More recently it
has been demonstrated that all human thyroid adenomas,
three types of human thyroid carcinoma tissues and mouse
non-cancerous follicular epithelial cells of the thyroid ex-
press RXFP2 mRNA (225). In addition, thyroid cancer cells
expressing INSL3 demonstrate enhanced motility and in-
creased colony formation in vitro and enhanced tumour
growth in vivo. Finally, the Ca
2
binding protein S100A4,
which increases cancer cell mobility and enhances tumour
tissue vascularization, is increased by INSL3, suggesting
that S100A4 is a downstream target of RXFP2 signaling in
thyroid cells (225).
RXFP2 mRNA expression is highest in rat kidney at late-
stage gestation and decreases dramatically at birth with the
lowest levels in adulthood (175). RXFP2 is expressed in
mesangial cells in mature glomeruli and inhibits prolifera-
tion of cultured primary glomerular cells, suggesting that
INSL3 and RXFP2 influence the genesis or maturation of
renal glomeruli and regulate mesangial cell density (175).
Pod1 is one of the main transcription factors involved in
glomerulogenesis (425), and an E-box consensus sequence
capable of binding Pod1 in conjunction with other helix
loop helix transcription factors is located upstream of the
gene for RXFP2 (176). Therefore, Pod1 may regulate ex-
pression of RXFP2 in the kidney during development to
facilitate glomerular cell proliferation (155). RXFP2 ex-
pression levels in the embryonic kidney were significantly
greater in Pod1 knockout mice than in heterozygous or
wild-type controls, indicating that RXFP2 is downstream of
Pod1 and Pod1 negatively regulates expression of RXFP2
in the glomeruli (155).
C. The Neuropeptide Relaxin-3 and Its
Cognate Receptor RXFP3
1. The relaxin-3 and RXFP3 system
The endogenous receptor for relaxin-3 is believed to be
RXFP3 on the basis of the coevolution of the ligand-recep-
tor pair (561), the overlap of relaxin-3 fibers, and RXFP3
expression in the brain (FIGURE 6) (46, 80, 314, 332, 334,
486, 487, 516) and the pharmacological profile of RXFP3
which shows high affinity for relaxin-3 (314, 545, 546).
However, at least pharmacologically, relaxin-3 may also
interact with RXFP1 (501) and RXFP4 (313) and may also
interact with RXFP2 in some species (453).
Relaxin-3 and RXFP3 expression is highest in the brain,
leading to a focus on roles of this ligand-receptor pair in the
CNS. Functional studies of relaxin-3-RXFP3 have been fa-
cilitated by the development of a chimeric peptide R3/I5
that activates RXFP3 and RXFP4 but not RXFP1 (312),
and R3(B23–27)R/I5, an antagonist with selectivity simi-
lar to the chimeric R3/I5 peptide (299) (see sect. IIC).
2. Role of the relaxin-3-RXFP3 system in stress
Relaxin-3 and RXFP3 are present in hypothalamic and ex-
trahypothalamic regions involved in the hypothalamic-pi-
tuitary-adrenal axis (312, 329, 487, 508). Corticotropin
releasing factor (CRF) is synthesized and released from the
paraventricular nucleus of the hypothalamus (PVN) in re-
sponse to stress. Interestingly, RXFP3 is abundantly ex-
pressed in the PVN, in addition to other regions commonly
associated with stress and anxiety, including the bed nu-
cleus of the stria terminalis, lateral septum, periaqueductal
gray, and the dorsal raphe (FIGURE 6) (312). The CRF
1
receptor mediates neural responses to CRF and most re-
laxin 3-containing neurons in the NI coexpress this receptor
(506, 516). The concept that relaxin-3 modulates stress
responses through interactions with CRF was investigated
by ICV administration of CRF to rats. This caused activa-
tion of relaxin-3 containing neurons, and neurogenic stres-
sors increased relaxin-3 mRNA in the NI (516). Further-
more, in NI neurons, relaxin-3 occurs in presynaptic vesi-
cles. This is in accord with the concept that relaxin-3
released from nerve endings in the NI is involved in regula-
tion of stress responses. Stress also transiently increases re-
laxin-3 mRNA in the rat brain following a forced swim
stress paradigm (25). These effects were partly blocked by
pretreatment with the CRF
1
antagonist antalarmin. While
these studies demonstrate the influence of CRF and stress on
relaxin-3, it is not yet clear whether relaxin-3 acting via
RXFP3 influences the CRF system.
3. Role of the relaxin-3-RXFP3 system in feeding and
metabolism
In the rodent, RXFP3 is expressed in the periventricular
hypothalamic nucleus, PVN, and supraoptic nucleus (SON)
(FIGURE 6) (486, 487), suggesting that the relaxin-3-RXFP3
system has a role in neuroendocrine and metabolic control.
Human relaxin-3 given ICV to rats either in the early light
or early dark phase, transiently but significantly increased
food intake (358). This effect did not result from increased
spontaneous activity or arousal. In addition, the orexigenic
effects of relaxin-3 are likely mediated by RXFP3, and not
RXFP1, since human relaxin-2 administration had no effect
(358).
The role of the PVN in this response was investigated using
acute and chronic human relaxin-3 infusions into the PVN,
both of which increased food intake (357). Human relax-
in-3 injection into the SON, arcuate nucleus (ARC), and
anterior preoptic area (APOA) also transiently increased
food intake. However, given that RXFP3 is not present in
the ARC or APOA, it is unclear how relaxin-3 exerts its
effects in these nuclei. Finally, ICV infusion of R3/I5 also
increased food intake, and this effect was blocked by pre-
administration of the antagonist R3(B23–27)R/I5 (299).
ICV administration of R3(B23–27)R/I5 alone has no ef-
fect on food intake (509), suggesting that under resting
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
453Physiol Rev VOL 93 JANUARY 2013 www.prv.org
conditions there is minimal tone in the relaxin-3-RXFP3
system.
Although relaxin-3 influences food intake, there is little ev-
idence to suggest that infusion of the peptide increases body
weight. In one study, chronic ICV human relaxin-3 did
increase body weight (219), but other studies have failed to
show weight changes following ICV or intrahypothalamic
injection of relaxin-3 (357, 358). In rats, chronic ICV
R3(B23–27)R/I5 did not affect body weight (509). There
is also limited evidence to suggest that body weight is influ-
enced by endogenous relaxin-3. Although relaxin-3 knock-
out mice were initially reported to have lower body weights,
this has not been replicated by other investigators (488,
509), suggesting that the decreased body weight may have
been due to differences in control animals, housing, or diet.
Alternatively, the chronic absence of relaxin-3 may be as-
sociated with compensation by other systems, whereas
acute infusions of relaxin-3 could transiently increase food
intake.
Several studies that investigated the effects of relaxin-3 in-
fusion on feeding also measured blood hormone levels.
Chronic ICV infusion of human relaxin-3 increased plasma
leptin and insulin levels (219). Similarly, increases in leptin
in ad libitum fed animals and TSH in ad libitum and pair fed
animals were observed with chronic infusion of human re-
laxin-3 into the PVN (357). These effects occurred in the
absence of changes in energy expenditure.
Studies of the roles of relaxin-3 in neuroendocrine function
and metabolism currently lack supporting mechanisms,
which prevents the generation of concrete conclusions.
More detailed studies are needed that examine the neural,
signaling, and gene transcription changes that lead to the
relaxin-3-induced feeding response. This may help to tease
out confounding effects on reward, motivation, and arousal
and determine whether there is a true feeding response to
relaxin-3 stimulation.
4. The relaxin-3-RXFP3 system in behavioral
activation and arousal
Much of the evidence for a role for relaxin-3 in behavioral
activation and arousal comes from the parallel study of the
neuroanatomy of relaxin-3 projections and sites of RXFP3
expression. Structures in the septohippocampal pathway of
rodents are heavily innervated by relaxin-3-positive projec-
tions from the NI (FIGURE 6). One of the major functions of
this pathway is the generation of hippocampal theta
rhythm, which has oscillations at 4–12 Hz. Theta rhythm is
controlled by pacemaker neurons of the medial septum
(MS) and is involved in behaviors such as vigilance, explo-
ration, orientation, navigation, locomotor control, and
working memory. The NI has an important role in theta
rhythm, and electrical stimulation of the nucleus causes
theta rhythm in the hippocampus. On the other hand, le-
sions of the NI disrupt theta rhythm initiated by stimulation
of the reticularis pontine oralis (390). RXFP3 signaling also
influences theta rhythm as in both anesthetized and con-
scious rats, RXFP3 modulates neuronal activity in the hip-
pocampus and MS to promote hippocampal theta rhythm
(330). In behavioral studies, blockade of RXFP3 in the MS
with R3(B23–27)R/I5 caused impaired performance in a
paradigm that investigated theta rhythm-dependent spatial
working memory. R3(B23–27)R/I5 produced dose-de-
pendent decreases in performance, which could be over-
come by coadministration of R3/I5 (330).
The relaxin-3-RXFP3 system also influences locomotor ac-
tivity in some models of rodent behavior. Chronic ICV in-
fusion of relaxin-3 had no effect on locomotor activity in
male Wistar rats (219), whereas ICV infusions of R3/I5
increased locomotor activity, and R3(B23–27)R/I5 had
no effect (509). However, female relaxin-3 knockout mice
were hypoactive in several paradigms including the locomo-
tor cell, large open field, Y maze, and novel object tests
(485). Whether these differences are due to gender or spe-
cies remains to be determined.
Serotonin (5-hydroxytryptamine, 5-HT) has well-estab-
lished roles in cognitive, emotional, and behavioral control
(reviewed in Ref. 106). Since the NI is located close to the
dorsal raphé, a region enriched in 5-HT neurons, studies
have been carried out examining the effects of 5-HT on
relaxin-3 expression (367) and showed that most relaxin-
3-containing neurons of the NI coexpress 5-HT
1A
recep-
tors. Inhibition of 5-HT synthesis for 3 days increased re-
laxin-3 mRNA in the NI. More studies are needed to deter-
mine the relationship between changes in 5-HT levels and
relaxin-3 expression.
5. Role of relaxin-3-RXFP3 in fluid homeostasis
Unlike the relaxin-RXFP1 system, the relaxin-3-RXFP3
system is not expected to influence fluid homeostasis. How-
ever, administration of human relaxin-3 into the lateral
ventricle causes drinking in rats, probably mediated by ac-
tions on RXFP1 in the SFO (45). Recently, it was demon-
strated that ICV relaxin-3 causes dose-dependent neuronal
activation in the hypothalamus and several circumventricu-
lar organs including the SFO, SON, and PVN (405). While
there is evidence for an interaction between relaxin-3 and
RXFP1 both in vitro and in vivo, there is no evidence that
this occurs physiologically, or that relaxin-3 has actions on
RXFP3 located in the circumventricular organs.
Thus the relaxin-3-RXFP3 system has a role in feeding,
metabolism, stress, arousal, learning, and memory. While
many of the hypotheses that have been developed on the
basis of neuroanatomical data have been tested in rodent
models, it is possible that the observed phenotypes do not
reflect the primary functions of the relaxin-3-RXFP3 system
in humans. Identification of the effector molecules down-
BATHGATE ET AL.
454 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
stream of RXFP3 and the development of alternative in vivo
models are important future steps in the validation of pu-
tative relaxin-3-RXFP3 functions.
D. INSL5 and RXFP4
RXFP4 (formerly GPCR142) was originally identified as a
receptor for relaxin-3 (313). However, the dramatically dif-
ferent expression patterns of relaxin-3 and RXFP4 sug-
gested that relaxin-3 was not the cognate ligand for RXFP4.
Furthermore, unlike relaxin-3, RXFP4 is less well con-
served and appears as a pseudogene in both rats and dogs,
suggesting that RXFP4 may be the receptor for a ligand
other than relaxin-3 in mammals. A related peptide, INSL5,
is also a pseudogene in rats and dogs, and its expression
pattern overlaps with RXFP4 leading to the suggestion that
INSL5 was the ligand for RXFP4. Subsequently, INSL5 was
demonstrated to be a high-affinity ligand for RXFP4 (315),
and mRNA is present in a variety of human tissues in-
cluding brain, kidney, testis, thymus, placenta, prostate,
salivary gland, thyroid, and colon (315), many of which
also coexpress INSL5. Overall, the receptor-ligand co-
evolution, pharmacology, and similarities in expression
profiles strongly indicate that RXFP4 is the endogenous
receptor for INSL5.
Transgenic INSL5 knockout/LacZ knockout mice devel-
oped to investigate the physiology of INSL5 (260) showed
-galactosidase staining of tissues that confirmed the distri-
bution of INSL5, with staining in the colon to a single
dispersed cell population. Mice overexpressing the INSL5
gene under control of the metallothionein or insulin pro-
moters have been produced, but there are no reports of any
physiological or behavioral findings related to INSL5 func-
tion.
E. Cross-reactivity Between Relaxin Family
Peptides and RXFP Receptors
1. Relaxin activates RXFP2 and RXFP3 in addition to
RXFP1
While RXFP1 is the cognate receptor for relaxin, some species
relaxins (including human relaxin-2) can also bind to and ac-
tivate the INSL3 receptor RXFP2. Human relaxin-2 and por-
cine relaxin bind to RXFP2 with nanomolar affinity and cause
cAMP accumulation (198, 199, 239, 301, 501). However,
while these relaxins bind to RXFP2, they do so with an
affinity lower than INSL3, and lower than that observed for
their interaction with RXFP1, consistent with the desig-
nated relaxin-RXFP1 and INSL3-RXFP2 ligand-receptor
pairings. Importantly both mouse and rat relaxin are unable
to bind to or activate RXFP2 in vitro (45, 454), and this
finding has been confirmed in vivo using transgenic animals
(65, 158, 274). This suggests that the human relaxin-2 in-
teraction with RXFP2 may be species specific (199).
While the original characterization of RXFP3 suggested a
unique relaxin-3-selective binding and activation profile
(314), a recent study also identifies binding and activation
of RXFP3 by relaxin (545). This may reflect the fact that
relaxin and INSL3 both retain homology within the domain
of relaxin-3 that dictates RXFP3 binding (299, 545) (see
sect. IIC). By examining a number of functional outputs
(ERK1/2 phosphorylation, cAMP inhibition, and NF
B
and AP-1 reporter gene activation), distinct patterns of cel-
lular responses to relaxin-3, relaxin, and INSL3 were iden-
tified (545). Relaxin-3 activated the widest range of signal-
ing pathways whereas relaxin and INSL3 activated only
subsets of these (see sect. VIC). Detailed binding studies
subsequently revealed that relaxin and INSL3 bind to a
distinct site on RXFP3, whereas relaxin-3 competes for
binding to RXFP3 with the
125
I-R3/I5 chimeric peptide,
relaxin-3, relaxin, and INSL3 could all compete for RXFP3
binding with
125
I-relaxin (545). Thus not only does relaxin
bind to and activate RXFP3 (in addition to RXFP1 and
RXFP2), but RXFP3 displays ligand-biased signaling, with
the pattern of cellular responses generated by this receptor
being highly dependent upon the ligand and on the cellular
background.
2. Relaxin-3 activates RXFP1, RXFP2, and RXFP4 in
addition to RXFP3
Prior to the discovery of a unique receptor for relaxin-3, it
was defined by its ability to bind to and activate RXFP1, but
with a lower affinity than relaxin (46). This property, in
conjunction with the inability of relaxin-3 to bind human
RXFP2, was subsequently exploited to describe a second
binding site for relaxin-3 within the second extracellular
loop of RXFP1 (501). This two-site binding model has since
been shown to be a general characteristic of relaxin family
peptide binding to both RXFP1 and RXFP2 (199, 210) (see
sect. IIIB). Although relaxin-3 does not bind and activate
human RXFP2, it has been demonstrated to weakly bind to
and activate rat RXFP2 (453).
The identification of RXFP3 as the cognate receptor for
relaxin-3 (314) also led to the subsequent deorphanization
of RXFP4 (the INSL5 receptor) (313), based on the high
degree of similarity between the two receptors and the abil-
ity of relaxin-3 to bind to and activate both receptors.
RXFP4 (human, monkey, bovine, porcine, and mouse) was
found to bind relaxin-3 with high affinity (98). Binding of
relaxin-3 to RXFP4 also caused cellular responses similar to
RXFP3, with inhibition of AC, increased GTP
S binding,
and increased intracellular calcium when RXFP4 is coex-
pressed with G
16
(98, 313, 315).
3. Interaction between INSL5 and RXFP3
Although RXFP4 was initially defined by its ability to bind
relaxin-3 (313), it was subsequently reclassified as the cog-
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
455Physiol Rev VOL 93 JANUARY 2013 www.prv.org
nate receptor for INSL5 (315). Based on the existing cross-
reactivity between many of the relaxin family peptides and
their receptors, the activity of INSL5 at all the relaxin fam-
ily receptors was subsequently assessed. Although INSL5
does not bind to RXFP1 or RXFP2, it does bind to RXFP3,
albeit with lower affinity than relaxin-3 (315). However,
INSL5 is an antagonist of responses to relaxin-3 (315). It is
not currently known if the peptide is a neutral antagonist,
inverse agonist, or biased-ligand. The lack of binding affin-
ity of INSL5 at RXFP1 or RXFP2 has been exploited in the
generation of the chimeric peptide (R3/I5) that is routinely
used to distinguish RXFP3 and RXFP4 binding sites from
those for RXFP1.
VIII. POTENTIAL THERAPEUTIC
APPLICATIONS FOR DRUGS
TARGETING RXFP RECEPTORS
A. RXFP1: A Promising Therapeutic Target
The long-established effects of relaxin on the adaptive
changes to pregnancy in women, in addition to its emerging
role in cardiovascular homeostasis, are the main targets of
current efforts in relaxin-RXFP1-based therapeutics. In ad-
dition, other physiological consequences of relaxin-RXFP1
interaction constitute additional, appealing therapeutic tar-
gets for future investigation.
1. Promotion of cervical ripening by relaxin
There is substantial experimental evidence for a role for
relaxin and RXFP1 in cervical ripening and parturition
(47). Based on this evidence, relaxin has been tested for its
ability to cause similar changes in women. Despite evidence
for cervical ripening in humans following application of
porcine relaxin (152, 339), there has been no demonstra-
tion of efficacy for human relaxin-2. Studies failed to show
efficacy of topically applied human relaxin-2, although this
may have been confounded by its low penetration into cer-
vical tissue (52). Thus, despite demonstrated efficacy in pro-
moting cervical ripening for human relaxin-2 application to
the common marmoset monkey (481), a similar role in
women is still uncertain. Despite this uncertainty, relaxin
may yet prove to be a useful adjunct in parturition, with
studies suggesting that both maternal and fetal relaxin may
contribute to rupture of the fetal membrane, particularly in
instances of premature birth (523).
2. Facilitation of implantation by relaxin
A principal role of relaxin, initially described in pregnant
guinea pigs (220), involves the preparation of the birth ca-
nal for parturition. Although the evidence for the hormonal
actions of relaxin is clear in rodents and some mammals, it
is of reduced importance in higher species. In women, max-
imum circulating levels of relaxin occur during the first
trimester of pregnancy (51, 138, 497), which is in direct
contrast to other species, such as rats (14, 108, 472) and
pigs (15, 53), where maximum circulating levels of relaxin
occur during the third trimester (393).
Due to the timing of this peak of circulating relaxin in
humans, the role of relaxin and RXFP1 in pregnancy is
more likely associated with first trimester physiological
events such as embryo implantation. Indeed, relaxin can
interact with RXFP1 expressed on endometrial and stromal
cells, inducing cAMP accumulation and the production of
vascular endothelial growth factor (VEGF) to cause decidu-
alization and preparation of the endometrium for implan-
tation (529, 540). More specifically, the identification of the
location and regulation of RXFP1 expression in the human
endometrium has supported a role for relaxin in the implan-
tation process: relaxin receptor binding in the uterus is dra-
matically increased just prior to, and peaks immediately
after, ovulation (67). This mirrored earlier studies monitor-
ing plasma relaxin levels in conceptive and nonconceptive
female cycles (497). Relaxin also promotes follicle develop-
ment in the human, as relaxin treatment of human ovarian
cortical tissue caused a significant increase in secondary
follicles and a decrease in primordial follicles, in addition to
causing increased vascularization of endometrial tissue and
thickening of the endometrium in preparation for implan-
tation (474). In the macaque, relaxin treatment has been
shown to assist in vitro fertilization due to improvements in
endometrial thickening and embryo implantation (214).
Thus the relaxin-RXFP1 system is an attractive candidate to
assist in implantation.
3. Relaxin treatment for the prevention of
preeclampsia
Preeclampsia commonly occurs during pregnancy and de-
scribes the reduced blood flow to the placenta leading to
systemic vasoconstriction, hypertension, renal dysfunction,
and ultimately multiple organ failure. Currently, symptoms
are usually relieved by preterm delivery of the fetus. Recent
studies suggest that the vasoconstriction that defines pre-
eclampsia may result from inhibition of factors such as
VEGF (308, 353, 593). There is preclinical evidence that
suggests that relaxin may be a useful treatment for pre-
eclampsia. In the clinical trials for the treatment of sclero-
derma, relaxin administration was associated with menor-
rhagia and metrorrhagia in a large number of women (540),
consistent with angiogenic and vasodilatory effects, and
likely due to increased VEGF expression and secretion in
human endometrial cells (408, 540). In the same study,
relaxin administration also decreased systolic and diastolic
blood pressure, consistent with a vasodilatory effect, and
improved renal function, suggesting that relaxin treatment
may relieve maternal symptoms of preeclampsia (538). Re-
laxin is currently being assessed as a treatment for pre-
BATHGATE ET AL.
456 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
eclampsia in a randomized, double-blind, placebo-con-
trolled clinical trial (538).
4. Relaxin as a novel treatment for congestive heart
failure
Relaxin is now emerging as a potential novel treatment for
acute congestive heart failure (526). The well-established
effects associated with relaxin during pregnancy including
increased cardiac output, blood vessel compliance, and re-
nal blood flow (265), together with studies that show pos-
itive inotropic and chronotropic effects on the heart (127,
271) and the observation that relaxin plasma levels increase
in heart failure (132), suggested that it may have beneficial
effects (reviewed in Ref. 135). In congestive heart failure,
the heart fails to pump enough blood to perfuse vital or-
gans. It is often associated with increased blood pressure,
and mortality is highly correlated with declining renal func-
tion and increased activity of the sympathetic nervous sys-
tem. Thus the ability of relaxin to induce vasodilation and
improve renal function in pregnancy, which was reflected in
the previous clinical trials for scleroderma (149, 527), has
suggested that the peptide may prove useful in the treatment
of heart failure.
In a phase I clinical trial examining the safety and dose-
response relationship of short-term infusion of relaxin on
hemodynamics in patients with congestive heart failure, ad-
ministration resulted in increased cardiac output and de-
creased systemic vascular resistance and pulmonary capil-
lary wedge pressure (133). Phase II clinical trials have since
been completed; the multicenter, placebo-controlled, dou-
ble-blind, randomized, international trial (Pre-RELAX-
AHF) assessed the effects of intravenous relaxin adminis-
tration in patients with acute heart failure (526). Despite no
significant improvement in renal function, relaxin was as-
sociated with a small but significant improvement in self-
reported breathlessness, and there was improvement in the
incidence of cardiovascular death and readmission (526).
Administration of the peptide was also associated with a
trend towards clinical improvement of heart failure and a
reduced requirement for diuretics. These studies again sup-
ported the safety of relaxin administration as a therapeutic
and confirmed the vasodilatory effects of relaxin in patients
with heart failure (133, 526). Thus there is promising scope
for the use of relaxin to improve patient outcomes following
congestive heart failure, especially in conditions of cardiac
fibrosis. The peptide is currently undergoing phase III clin-
ical trials (RELAX-AHF-1) (421).
5. Antihypertensive effects of relaxin
Relaxin also exerts a local influence on the circulation by
increasing vasodilation and passive compliance: both re-
laxin and RXFP1 mRNA have been identified in thoracic
aortas, small renal arteries, and mesenteric arteries in both
rats and mice (388). Several studies suggest that relaxin
exerts these local effects through the synthesis and produc-
tion of NO, particularly in situations of natural or induced
myocardial infarction (33, 154, 343, 346). This is sup-
ported by studies in the relaxin-knockout mouse, which
demonstrated blunted NO-dependent myogenic reactivity
(388). Relaxin also induces vasodilation in human systemic
resistance arteries in vitro (170). The vasodilation of gluteal
resistance arteries was endothelium dependent (170) and
substantially inhibited in arteries from patients being
treated with ACE inhibitors. Inhibition of cyclooxygenase
(indomethacin) or soluble guanylyl cyclase (ODQ-4M) also
greatly reduced the vasodilator response to relaxin (171). In
addition to these direct vasodilator effects of relaxin, recent
evidence also suggests that relaxin treatment reverses the
arterial remodeling that is a characteristic feature in spon-
taneously hypertensive rats with age and improves large
artery compliance (568). Thus the actions of relaxin on
connective tissue metabolism may also contribute to the
long-lasting effects of a relatively short period of relaxin
treatment.
In clinical studies assessing the potential efficacy of relaxin
in the treatment of scleroderma, the peptide caused a signif-
icant decrease in both systolic and diastolic blood pressure,
that was consistent with vasodilation (528). Clinical trials
investigating relaxin as a treatment for congestive heart
failure also demonstrated decreased blood pressure (134,
526).
6. Antifibrotic effects of relaxin
One of the physiological effects of relaxin with broad ther-
apeutic potential is its effects on connective tissue regula-
tion and fibrosis. From a therapeutic viewpoint, relaxin
decreases excess collagen deposition in fibrotic lesions, with
a conservation of endogenous connective tissue structure
(181, 542). Relaxin also exerts anti-inflammatory effects,
can inhibit the activation of human neutrophils by pro-
inflammatory agents (345), and prevents histamine and
granule release by activated basophils (29) and mast cells
(344). In an animal model of allergic airway disease, there
was increased collagen deposition and decreased expression
of MMP9 in relaxin-knockout animals compared with con-
trols (371), although these studies demonstrated receptor
involvement only in airway fibrosis homeostasis and not
during the inflammation-induced fibrosis associated with
chronic allergic airway disease (441). Taken together, re-
laxin is protective during the allergic inflammatory re-
sponse. In rodent models of angiogenesis and wound heal-
ing, relaxin caused increased vascularization and neoangio-
genesis of ischemic wound sites (309, 541) but had no effect
on cytokine expression in cells taken from nonwound sites
(541), indicating promising specificity for therapeutic appli-
cations.
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
457Physiol Rev VOL 93 JANUARY 2013 www.prv.org
7. Roles of relaxin in cancer
As discussed in the previous section, relaxin appears to be
recruited in many types of cancer cells as an endogenous
factor for tissue remodeling. In particular, relaxin is associ-
ated with prostate cancer progression, and increased ex-
pression of relaxin (but not RXFP1) occurs in prostate car-
cinomas and prostate cancer cell lines (157, 532, 549). Im-
portantly, downregulation of either relaxin or RXFP1
caused significant inhibition of growth (549) and invasive-
ness, in addition to increased apoptosis (157) in rodent
prostate cancer models. Furthermore, lentiviral driven ex-
pression of an RXPF1 peptide antagonist within prostate
tumors in a mouse model of human prostate cancer sup-
pressed tumour growth, whereas control lentiviruses had
no effect (480). Hence, drugs that block relaxin action in
prostate cancer may be an effective means to slow prostate
cancer progression. Furthermore, relaxin is associated with
increased invasiveness of endometrial (273), breast (59,
60), and thyroid (226) carcinomas, suggesting that agents
blocking relaxin action may be used for the possible treat-
ment of multiple cancers.
B. RXFP2 and Its Potential as a Therapeutic
Target
INSL3 and its cognate receptor RXFP2 are clearly key reg-
ulators of the reproductive system in both male and female
mammals. Apart from reproductive tissues, thyroid and
bone are the only other major tissues to express the peptide
and receptor. There is also good evidence for RXFP2 ex-
pression in brain, but no studies to date have provided
evidence for the expression of INSL3 in either laboratory
animals or humans.
1. RXFP2 as a target for modulation of fertility
The major physiological effects of the INSL3/RXFP2 sys-
tem are on the reproductive system: lack of either INSL3 or
RXFP2 in males results in cryptorchidism (379, 596), and
overexpression of INSL3 in females causes ovarian descent
and bilateral inguinal hernia (490). INSL3 stimulation of
RXFP2 in germ cells causes meiotic progression of arrested
oocytes in preovulatory follicles and suppresses male germ
cell apoptosis (276). However, it is likely that targeting
RXFP2 in the gubernaculum in utero would not be a useful
therapeutic strategy, and indeed, most male babies with
cryptorchidism are treated surgically. However, modula-
tion of fertility is an attractive therapeutic option for the
INSL3-RXFP2 system. Activation of RXFP2 expressed on
germ cells suppresses male germ cell apoptosis (276),
whereas injection into rat testis of an INSL3-based antago-
nist targeting RXFP2 led to a significant decrease in testis
weight, consistent with inhibition of spermatogenesis
(121). Furthermore, in a study examining a male hormonal
contraceptive regime (involving administration of testoster-
one with a progestagen over 6 mo), those patients with
persistent sperm production had higher serum INSL3 levels,
suggesting again that the INSL3-RXFP2 system plays a role
in spermatogenesis (8). Thus it is tempting to speculate that
an INSL3-based antagonist targeting RXFP2 could be an
effective male contraceptive agent. Additionally, INSL3 in
females is associated with oocyte maturation and follicular
development, and therefore, INSL3 may be effective for
certain types of female infertility.
2. Other potential therapeutic applications of RXFP2
Interestingly, in a manner similar to relaxin, the INSL3-
RXFP2 system has been shown to be present in human
prostate carcinoma cell lines and human thyroid carcino-
mas, and treatment of these cell lines with INSL3 caused
increased tumor cell motility (225, 286, 429). Another
study suggests that INSL3 and RXFP2 may also have a role
in osteoporosis (162). Thus there may be further therapeu-
tic potential for this ligand-receptor system in pathologies
including cancer and osteoporosis in the future.
C. RXFP3 as a Target for the Control of
Anxiety and Food Intake
The development of therapeutics based on the relaxin-3-
RXFP3 system is currently at the proof of concept stage.
Much of what we know about the functions of the relaxin-
3-RXFP3 system is based on a small number of anatomical,
pharmacological, and genetic interference studies in rodent
models. In addition, expression of RXFP3 has only been
detected in human tissue by RT-PCR. However, relaxin-3
has recently been detected in the macaque by both in situ
hybridization and immunohistochemistry, suggesting that
the relaxin-3-RXFP3 system may also be present in humans
(332).
Nonetheless, there is substantial interest in the therapeutic po-
tential of the relaxin-3-RXFP3 system. In general, GPCRs are
attractive drug targets due to cell surface expression and
thus good drug accessibility (particularly hydrophilic
drugs) and selectivity due to the distribution of expression
in different tissues and cell types. In addition, GPCRs are
often responsible for maintaining physiological homeosta-
sis by modulating the actions of neuropeptides and neu-
rotransmitters. As the receptor of a neuropeptide, RXFP3 is
particularly appealing. Many neuropeptides are colocated
with neurotransmitters and act synergistically with them to
modulate their actions. Thus neuropeptide receptors pro-
vide a unique opportunity to influence systems with less
dramatic consequences than interfering with amino acid
transmitters themselves. Since relaxin-3 is produced in
GABAergic neurons of the nucleus incertus, it likely acts in
concert with GABA signaling. In this way, it may be possi-
ble to fine-tune the GABAergic system by targeting RXFP3.
This section focuses on the physiological functions of the
BATHGATE ET AL.
458 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
relaxin-3 and RXFP3 system and how this may be manip-
ulated for therapeutic purposes.
1. The relaxin-3-RXFP3 system as a target for
antianxiety drugs
The relaxin-3-RXFP3 system is clearly implicated in the
regulation of the stress response. As previously mentioned
(see sect. VIIC), the great majority of relaxin 3-containing
neurons in the NI express CRF-R1 (422). Additionally, the
principal site of CRF production in the brain, the PVN, also
expresses RXFP3 (516). In rats, relaxin-3 neurons are acti-
vated by CRF administration, and expression of relaxin-3
in the NI is increased in response to swim stress (516).
Similarly, repeat forced swim has been demonstrated to
increase relaxin-3 expression, while pretreatment with the
CRF-RI antagonist antalarmin attenuated the response
(25). Currently, the physiological response to elevated re-
laxin-3 expression, the signals necessary to stimulate relax-
in-3 release, and the influence of stress on RXFP3 signaling
are unknown. From neuroanatomical data, it appears that
RXFP3 could be part of a feedback system that regulates
continued CRF release following stress; however, this has
yet to be demonstrated experimentally. While these ques-
tions remain, the potential therapeutic application of
RXFP3 modulators for the stress system remains unre-
alised.
2. Modulation of feeding by the relaxin-3-RXFP3
system
A role for the relaxin-3-RXFP3 system in energy homeosta-
sis has been suggested by studies of peptide infusion into the
brain, which demonstrate effects of relaxin-3 on body
weight, food intake, and plasma hormone levels. Chronic
ICV human relaxin-3 has been reported to increase body
weight (219), although other studies showed no change in
body weight after ICV or intrahypothalamic nuclei injec-
tion of relaxin-3 (357, 358).
Although the evidence for a role for relaxin-3 in body
weight control is equivocal, there are indications that the
peptide may influence food intake. Initially this was
prompted by the identification of RXFP3 expression in the
hypothalamus (312, 329, 508). In this context, human re-
laxin-3 administered ICV to rats either in the early light or
early dark phase transiently increased food intake (219),
and acute or chronic human relaxin-3 infusion into the
PVN also increased food intake (357). Other hypothalamic
nuclei, such as the SON, ARC, and APOA, may also be
involved, as food intake was transiently increased by hu-
man relaxin-3 injection into these regions. However, given
that RXFP3 is not present in the ARC or APOA, it is unclear
how relaxin-3 exerts its effects at these sites. However, ICV
infusion of the RXFP3 agonist R3/I5 also increased food
intake, and this effect was blocked by preadministration of
the RXFP3 antagonist R3(B23–27)R/I5 (299). Interest-
ingly, chronic ICV administration of R3(B23–27)R/I5 had
no effect on feeding, but significantly increased body weight
(509). Studies in relaxin-3 knockout mice have provided
equivocal data for a role of the relaxin-3-RXFP3 system in
body weight control, with one study describing a smaller
leaner phenotype (509) whereas another (using the same
knockout mice) failed to demonstrate any differences (485).
Several studies that investigated effects of relaxin-3 infusion
on feeding also measured blood hormone levels. Chronic
ICV human relaxin-3 increased plasma leptin and insulin
levels (219), and chronic infusion of human relaxin-3 into
the PVN increased leptin and decreased TSH in ad libitum-
fed rats (357). These effects were observed in the absence of
changes in energy expenditure. The levels of plasma endo-
crine factors were also measured in the chronic RXFP3
agonist and antagonist studies conducted in rats. Chronic
administration of the RXFP3 agonist R3/I5 significantly
increased insulin, leptin, adiponectin, testosterone, and an-
giotensinogen levels and decreased growth hormone levels
(509). Infusion of the RXFP3 antagonist R3(B23–27)R/I5
also decreased growth hormone levels. Relaxin-3 knockout
mice at 10–15 wk have reduced leptin and at 30–35 wk
reduced insulin and leptin (509).
Given the lack of consistent evidence and the numerous
hypothalamic neurotransmitter and peptidergic systems
that regulate energy balance, it is difficult to assess the po-
tential of the relaxin-3-RXFP3 system as a drug target. Al-
though relaxin-3 administration increases food intake in
rodents, this effect is also seen in regions lacking RXFP3,
questioning its specificity. Infusions of R3(B23–27)R/I5
are only effective in inhibiting relaxin-3 stimulated, but not
basal food intake. The results suggest that the complex
regulation of energy homeostasis by the CNS is difficult to
override by manipulation of a single neuropeptide/receptor
system. Nevertheless, evidence from rodent models suggests
that RXFP3 agonists and antagonists may influence energy
homeostasis in a manner that may be useful for the treat-
ment of human metabolic conditions such as anorexia, ca-
chexia, and obesity.
IX. MAJOR UNANSWERED QUESTIONS
AND FUTURE DIRECTIONS
A. Signaling Paradigms
1. Role of signalosomes in relaxin family peptide
receptor signaling
The recent demonstration that the expression of RXFP1
receptors in cells induces the formation of signalosomes
that are constitutively active and highly sensitive to attomo-
lar concentrations of relaxin adds a new dimension to re-
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
459Physiol Rev VOL 93 JANUARY 2013 www.prv.org
laxin signaling (see sect. VIA). A particularly interesting
feature of the RXFP1-signalosome is that attomolar con-
centrations of relaxin stimulate small increases in cAMP,
whereas nanomolar concentrations preferentially activate
the classical pathways involving G
s
,G
oB
, and G
i3
(FIG-
URE 7A). Importantly, these pathways appear to act quite
independently, and there is no effect of inhibitors of classi-
cal pathway-specific proteins (including G
i/o
, PI3K and
PKC) on cAMP generated in response to subpicomolar con-
centrations of relaxin. Equally there is no effect of inhibition of
signalosome-specific proteins (including AC2, AKAP79, and
-arrestin 2) upon classical relaxin cAMP signaling (FIGURE
7A). Furthermore, higher concentrations of relaxin are as-
sociated with dissociation of the signalosome, which is fol-
lowed by activation of the classical signaling pathways. The
pathways also generate cAMP in quite distinct regions of
the cell. Thus signalosome-specific AC2 is known to be
preferentially excluded from lipid-rich domains (565),
whereas activation of the G
i3
pathway with nanomolar
relaxin concentrations depends on lipid-rich domains in
HEK293 cells (204). Since AC2 expression occurs predom-
inantly in brain, lung, skeletal muscle, heart, and uterine
myometrium (438, 565), it is likely that there are both tis-
sues that display RXFP1 signalosome signaling and those
that do not which may help to determine the physiological
role of signalosome-localized RXFP1. It is interesting to
note that previous studies of the regulation of RXFP1 ex-
pression by estrogens showed marked differences between
tissues, and receptor expression was markedly upregulated
in uterus but not affected in brain or heart (515). The scaf-
folding of AC2 to RXFP1 is highly dependent on AKAP79
that is known to scaffold many other proteins including
other GPCRs, PKC, several ion channels, and a number of
AC isoforms (see Ref. 196). The activity of AC isoforms
may be variably influenced by interaction with AKAP79
and AC2 activity is inhibited (142), although this appears to
be primarily offset in RXFP1 signalosomes by the scaffold-
ing of RXFP1 and AC2. The regulatory complex of RXFP1,
-arrestin-2, and PDE4D3 also has some unusual features
that call into question some of the current paradigms con-
cerning the role of
-arrestins in receptor desensitization
and internalization. While the
2
-adrenoceptor is phos-
phorylated by GRK-2 following receptor activation and
subsequently becomes a substrate for
-arrestin-2 which
triggers formation of clathrin-coated vesicles and internal-
ization (for review, see Ref. 328),
-arrestin-2 clearly has a
different role with RXFP1. The interaction between RXFP1
and
-arrestin-2 is constitutive and does not involve recep-
tor activation or phosphorylation and does not appear to be
involved in desensitization (201). Although this interaction
involves S704 of RXFP1 (201), this residue does not appear
to be phosphorylated, and indeed, exposure of RXFP1 to
higher concentrations of relaxin fails to cause phosphory-
lation, desensitization, or internalization of the receptor
(87, 514). It seems likely that in this case
-arrestin-2 acts
solely as a scaffold for the formation of the regulatory com-
plex and has no role in receptor desensitization or internal-
ization.
It is interesting to speculate whether the constitutive activity
of the RXFP1 signalosome complex provides signaling that
regulates expression of proteins necessary for signalosome
formation or indeed the conventional relaxin signaling
pathways. It will also be of interest to determine whether it
is the specific activation of lipid raft-based signaling path-
ways that provides the stimulus for the dissociation of the
RXFP1 signalosomes. More overt physiological applica-
tions of this highly sensitive, concentration-dependent sig-
nalosome are not yet clear. However, it is well recognized in
humans that relaxin has a role during embryo implantation
where it is required to exert its effects over an extended
period of time, exemplified by endometrial decidualization
which is crucial for embryo implantation and the mainte-
nance of pregnancy. It has been demonstrated that endome-
trial decidualization is dependent on both circulating and
locally produced relaxin (146, 214, 409, 540) and requires
the maintained elevation of cAMP (reviewed in Ref. 529).
Thus signalosome relaxin signaling may provide a mecha-
nism that provides a physiological, perhaps homeostatic,
role for the low concentrations of peptide present in the
circulation, and thus may be linked to some of the long-
term physiological effects. Future research, perhaps utiliz-
ing relaxin knockout models or single-cell models of phys-
iologically relevant targets (i.e., atrial cardiomyocytes, en-
dometrial cells, fibroblasts, or neurons), will be required to
fully elucidate the downstream consequences and physio-
logical significance of this ultrasensitive cAMP signaling
pathway.
2. Role of LDLa modules in RXFP1 and RXFP2
signaling
Relaxin family peptides interact with RXFP1 and RXFP2 in
a complex manner (see sect. VI, Aand ). Not only do the
peptides bind to the receptors in the LRR region and in
ECL2, but also activation of signaling requires the presence
of an intact LDLa module at the NH
2
terminus. A splice
variant of RXFP2 that lacks the LDLa module or an engi-
neered RXFP1 equivalent has binding properties similar to
the intact receptor but do not produce a cAMP response to
either INSL3 or relaxin, respectively (456). Also, the addi-
tion of a soluble form of RXFP1 LDLa to cells expressing
RXFP1 inhibits the cAMP response to relaxin. Thus the
activation of the cAMP signaling pathway by relaxin in-
volves at least three events: high-affinity binding to the LRR
region, binding to a secondary site on ECL2, and the in-
volvement of the LDLa module. The precise mechanism
whereby the LDLa module activates signaling is not known,
although several features of the region are essential for ac-
tivity. The RXFP1 LDLa has three disulfide bonds formed
by six conserved cysteines and mutations of these bonds
cause loss of the cAMP response. These three disulfide
bonds are thought to provide the structural integrity neces-
BATHGATE ET AL.
460 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
sary for four acidic residues (N26, D30, D36, and E37) to
form a correctly folded cage that ligates a Ca
2
essential for
activity (231). However, other specific residues are also nec-
essary for activity, since swapping of the RXFP1 LDLa for
the LB2 domain of the LDLR that has high structural and
sequence similarity does not produce a functional receptor.
Site-directed mutagenesis of RXFP1 has identified L7, Y9,
L22, and L23 as residues that influence the activity of the
LDLa region, and mutations of these residues generate re-
ceptors that display reduced potency in response to relaxin.
Some gain-of-function studies have been completed that
indicate that substitution of the first nine residues of LB2
with those of the RXFP1 LDLa restores some cAMP re-
sponsiveness; however, the profile required to restore full
activity has yet to be determined.
All of the current studies have examined whether particular
structural features of the LDLa module are associated with
cAMP production, but perhaps given the pleiotropic nature
of RXFP1 signaling, it would also be worth examining if all
signaling pathways are affected. Conceivably the confor-
mations of RXFP1 associated with cAMP signaling may
require the LDLa contribution but could differ from con-
formations that result in the activation of other pathways.
Several models have been proposed to explain the role of
the LDLa module in signaling. One involves a receptor
monomer with the LDLa module interacting with the TM
regions of the receptor to induce conformations that pro-
mote efficient interaction with G proteins and signaling. In
this and alternative models, the LDLa acts as a tethered
ligand. An additional model suggests that the LDLa may
have a role in the formation of receptor dimers and in this
case interacts with the dimerization partner. Both models at
present lack substantial evidence to support or challenge
their credibility. What is known is that the LDLa modules
on their own or in combination with an LDLa-less receptor
do not signal, so the link between the LDLa module and the
receptor is essential.
3. Dimerization of relaxin family peptide receptors:
does it occur and what is the functional role?
There is evidence to suggest that both RXFP1 (282, 511)
and RXFP2 (510) form homodimers and heterodimers
(510). The BRET signal is saturable but is not affected by
the addition of the cognate ligand (human relaxin-2 for
RXFP1; INSL3 for RXFP2). It is likely that dimerization
primarily involves the 7-TM region of RXFP2 since a con-
struct (RXFP2–TM1–7–GFP
2
) that lacks the extracellular
domain of the receptor and hence the primary ligand bind-
ing site still forms dimers with RXFP2-RLuc. It has been
suggested that dimer formation is necessary for signal trans-
duction with ligand binding occurring at the LRR region of
one dimer partner followed by interaction of the bound
ligand with the ECL2 of the second partner and initiation of
signaling (510). However, at present, there are no experi-
ments with two inactive mutant receptor dimer partners
that show complementation and demonstrate rescue of
function by dimerization to support this concept. It has also
been suggested that dimer formation may explain apparent
negative cooperativity observed with both RXFP1 and
RXFP2. With both receptors, the rate of dissociation fol-
lowing incubation of cells expressing either receptor with
either radiolabeled human relaxin-2 or INSL3 was deter-
mined after allowing the system to equilibrate and then
undergo infinite dilution in the absence or presence of un-
labeled ligand. The presence of unlabeled ligand was asso-
ciated with modest increases in the rate of dissociation,
suggesting negative cooperativity where the binding of li-
gand to each receptor is associated with a decrease in affin-
ity of the remaining unoccupied receptors. It should be
borne in mind, however, that such behavior does not nec-
essarily reflect the formation of dimers (93), and studies are
required utilizing receptors that retain function but are un-
able to form dimers or on receptors expressed in model
phospholipid bilayers that allow examination of their func-
tional characteristics when in monomeric form (548, 559).
4. Bell-shaped concentration-response relationships:
explanation and relevance to clinical actions
A commonly observed characteristic of concentration-re-
sponse curves to relaxin mediated by RXFP1 is that they are
bell-shaped. These include cAMP responses in HEK293
cells transfected with RXFP1 (198); endometrial glandular
epithelial cells (96); increases in L-type Ca
2
current in
rabbit sinoatrial node cells (206); MMP1 expression in hu-
man lung fibroblasts (542); changes in glomerular filtration
rate, renal plasma flow, renal vascular resistance, and
plasma osmolality (112); as well as effects on dyspnea and
survival in the phase II clinical trial of relaxin in cardiac
failure (526). Similar findings have been reported for
RXFP2 for cAMP accumulation (198) and negative coop-
erativity (510).
Although the phenomenon of bell-shaped concentration-
response curves has been well described, there are no clear
explanations, even though the clinical findings suggest that
a better understanding of the process will be necessary to
gain the optimum therapeutic benefit from relaxin treat-
ment. One mechanism that has been suggested to explain
the curves is negative cooperativity associated with receptor
dimerization (510, 511). This would argue that increases in
ligand concentration and receptor occupancy would reduce
the affinity of the receptor for further interactions. How-
ever, one might expect in these circumstances that RXFP1
and RXFP2, receptors that both display negative coopera-
tivity, would exhibit bell-shaped concentration-response
curves. However, this does not appear to be the case as
cAMP accumulation in response to activation of RXFP1 at
3 or 30 min or RXFP2 at 30 min display bell-shaped curves,
whereas RXFP2 at 3 min does not (198). It could be argued
that the RXFP2 system at 3 min is not in equilibrium and
the effect of increasing receptor occupation balances out the
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
461Physiol Rev VOL 93 JANUARY 2013 www.prv.org
reduced receptor affinity. However, this does not convinc-
ingly explain the difference between RXFP1 and RXFP2 at
the same time point. The different patterns are unlikely to
be related to a different time course of receptor coupling to
G proteins as this occurs rapidly after receptor activation
(198). Alternatively, the decrease in response with higher
concentrations of peptide could be due to desensitization
and/or internalization of the receptor. For RXFP1, this is an
unlikely explanation. Even in early functional experiments
examining relaxin responses in target tissues such as the
heart, the maximum responses were maintained for over 6 h
even with regular washing of the tissue (514), and more
recent studies showed no significant receptor phosphoryla-
tion,
-arrestin recruitment, or internalization following
exposure to ligand (87). Similar findings have been reported
in cell-population and single-cell studies (196). Even in
studies where desensitization of RXFP1 to overexpression
of
-arrestin-2 has been demonstrated, the effects are rela-
tively small and associated with a small degree of internal-
ization (281). Recent studies from our own lab directly
examining the interaction of RXFP1 and
-arrestin-1 or
-arrestin-2 using BRET provided no support for interac-
tion between these proteins (Kocan, unpublished observa-
tions). On balance, it seems unlikely that internalization
and desensitization can account for the phenomenon. An-
other possibility reflects the multiple G proteins known to
interact with RXFP1 and RXFP2. Both receptors interact
with G
s
and G
OB
, and in addition, RXFP1 also recruits
G
i3
. If the efficiency of coupling of G
s
and G
i3
were
high, then the usual concentration-response curve would be
generated with low concentrations of ligand. If this is fol-
lowed however, by lower efficiency coupling to G
OB
which inhibits AC, then inhibition of cAMP production
could occur with higher concentrations of peptide. This,
however, cannot explain the lack of a bell-shaped curve for
RXFP2 stimulation at 3 min, and in addition, this type of
behavior would be unlikely for signaling pathways that do
not involve activation of AC. Finally, the effect could be due
to interaction between the numerous signaling pathways
activated by relaxin family peptides. For instance, another
G protein-coupled receptor, the
3
-AR, shows clear evi-
dence of interaction between cAMP and MAPK signaling.
The
3
-AR, like RXFP1 and RXFP2, activates multiple sig-
naling pathways, and one striking finding is that an antag-
onist of cAMP signaling is a strong activator of p38 MAPK
signaling (449). However, while the cAMP response to ago-
nists is increased with increased receptor expression, the
p38 MAPK response is almost completely abolished. This
suggested that increased cAMP generation associated with
increased receptor expression might be inhibiting p38
MAPK activity, and this was supported by inhibition of the
MAPK response by a cell-permeable cAMP analog (449).
As indicated earlier (see sect. VIA), both RXFP1 and
RXFP2 couple to G
s
to increase cAMP and to G
oB
to
negatively modulate this effect (198). Only RXFP1 couples
to G
i3
to cause a further surge of cAMP by releasing G
␤␥
subunits that activate PI3K to promote PKC-
transloca-
tion to the plasma membrane where it stimulates AC5
(198, 204, 382, 383). The translocation step is believed
to be responsible for the delay in the G
i3
-mediated surge
in cAMP accumulation. However, the G protein activa-
tion and release of G
␤␥
occurs rapidly after receptor
activation (198), and in addition to effects on PI3K, G
␤␥
subunits can also directly influence AC activity (200).
Thus the different pattern of appearance of G
␤␥
subunits
following activation of RXFP1 and RXFP2 could influ-
ence the concentration-response relationships observed
at early stages following receptor activation.
5. Ligand-directed signaling bias: a common
phenomenon in relaxin family peptide receptors or
an idiosyncrasy of RXFP3?
Ligand-directed signaling bias describes the selective stabi-
lization of particular receptor conformations by ligands re-
sulting in selective activation of downstream signal trans-
duction pathways (22, 151, 277, 278). Although human
relaxin-3 or human relaxin-3 -chains were originally
thought to be the only ligands that interacted with RXFP3,
there is now evidence that several relaxin peptides interact
with RXFP3 to activate distinct signaling profiles through
different, although sometimes overlapping pathways (545).
Recently, these studies have been extended to examine the
specific signaling pathways activated by human relaxin-2
and human relaxin-3 and also to examine whether the
RXFP3 antagonist R3(B23–27)R/I5 displayed signaling
bias. Human relaxin-2 was found to activate p38MAPK
and ERK1/2 with lower efficacy than human relaxin-3, but
both peptides have similar efficacy for JNK1/2/3 phosphor-
ylation and the responses are PTX sensitive. In contrast,
R3(B23–27)R/I5 activated ERK1/2 in a PTX-sensitive
manner, p38MAPK in a PTX-insensitive manner, and did
not activate JNK. Antagonism by R3(B23–27)R/I5 was
only seen for responses mediated by human relaxin-3 (Ko-
can et al., unpublished data).
More recently, we have examined a wide range of ligands
known to interact with RXFP1. These varied in potency
and efficacy across signaling pathways including cAMP ac-
cumulation, ERK1/2, and p38MAPK phosphorylation, but
there was no indication of ligand-directed signaling bias
with agonists, antagonists, and related peptides such as
INSL3 and INSL5 (Siwek, Kocan, and Summers, unpub-
lished data). It is possible that the lack of signaling bias
observed with the ligands so far tested relates to the mode of
action of RXFP1 where the interaction of the peptide ligand
with the LRR region and the ECL region causes the LDLa
“tethered” ligand to interact with the receptor. Since the
final common pathway is the same for all ligands, this may
explain the inability of the receptor to adopt alternative
signaling conformations. This may become possible if li-
gands that can be developed that target the site on the
receptor utilized by the LDLa module. Alternatively, li-
BATHGATE ET AL.
462 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
gands with an allosteric mode of action may be able to
promote signaling bias. Likewise, no studies to date have
reported signaling bias at RXFP2 and RXFP4. However, it
should be borne in mind that studies to date with relaxin
family peptide receptors have largely been carried out with
a comparatively restricted range of peptides. Only recently
has evidence emerged for short linear peptides that can
activate RXFP1 and RXFP2 (see sect. IXD) (466, 467).
B. Receptor Roles and Potential
1. RXFP2 but not INSL3 in brain
One apparent anomaly that has been described for RXFP2
is the presence of the receptor but not its cognate ligand in
brain. The evidence for the presence of the receptor includes
RT-PCR, in situ hybridization histochemistry, and direct
binding studies with
125
I-INSL3 (239, 460, 469). In con-
trast, there is no convincing evidence for the expression of
INSL3 in the brain of mammals with the exception of bo-
vine hypothalamus, perhaps reflecting the lack of a relaxin
system in ruminants. Explanations that have been put for-
ward include that brain RXFP2 remains as part of a now
defunct ligand-receptor pairing (561) or that it is the target
for circulating INSL3 that either is transported across or
acts on receptors outside the blood-brain barrier. Sugges-
tions that relaxin might act as a ligand for RXFP2 in brain
appear unlikely due to lack of colocalization.
2. RXFP3: does it have a role in feeding/weight
control or is it one of many systems controlling
these functions?
The relaxin-3-RXFP3 ligand-receptor pairing is predomi-
nantly a neuropeptide receptor system. The cell bodies are
found mainly in the NI and project to many areas of the
brain. There has been a great deal of interest in this system
because of its reported roles in stress and feeding. Most
neurons that express relaxin-3 in the NI also express CRF-
R1, the receptor that responds to CRF released by stress.
Administration of CRF increases c-fos expression in many
relaxin-3 expressing neurons, and swim stress increases re-
laxin-3 mRNA in the NI (86). Although these interactions
are of interest, more work needs to be done to establish the
relationship between stress and the relaxin-3 system.
The interest in the role of relaxin-3-RXFP3 in feeding was
stimulated by the presence of RXFP3 in the hypothalamic
paraventricular nuclei. ICV injection of relaxin-3 or the
RXFP3 agonist R3/I5 into the brain of rats increased feed-
ing, and the effects are blocked by the antagonist R3(B23–
27)R/I5 (486). Human relaxin-2 given in the same manner
decreased food intake, perhaps reflecting its actions on
RXFP1 or alternatively its different spectrum of action at
RXFP3 (545). Exploitation of the contrasting effects of bi-
ased agonists at RXFP3 would establish a novel paradigm
for drug action. However, the actions of relaxin-3 on feed-
ing do not appear to involve changes in expression of other
hypothalamic peptides known to have roles in appetite con-
trol including neuropeptide Y, proopiomelanocortin, or ag-
outi-related peptide (486). It has been suggested that the
actions of relaxin-3 on feeding do not represent a primary
role for this neuropeptide system but rather reflect its wider
role as an arousal neuromodulator (486).
3. What is the role of RXFP4?
Studies of the physiological role of RXFP4 are at a much
earlier stage than those of RXFP3. Although human relax-
in-3 was found to activate RXFP4 (GPR142) (313), it was
quickly recognized that this peptide was unlikely to be the
cognate ligand (98) based largely on the tissue expression
profile. Whereas relaxin-3 is largely confined to the brain,
RXFP4 has a wide tissue distribution including colon, kid-
ney, placenta, prostate, thymus, thyroid, and salivary gland
with only small amounts found in brain (313). Subsequent
work provided strong evidence that a related peptide INSL5
was likely the cognate ligand for RXFP4, since it showed a
similar tissue distribution and selectively activated the re-
ceptor (315). Although RXFP4 in the rat is a pseudogene
and in mice shows some sequence variation, in other species
such as human, monkey, cow, and pig, it is highly conserved
and displays similar ligand specificity (98). At the present
time, the physiological role of INSL5-RXFP4 is unclear,
although a recent report suggests that polymorphisms of
RXFP4 may be associated with obesity (376). This is of
particular interest given the potential role of relaxin-3-
RXFP3 in appetite control (see NOTE ADDED IN PROOF 1).
C. Therapeutic Actions of Relaxin
1. What is the mechanism of action of relaxin in the
treatment of cardiac failure?
Cardiac failure is a complicated disease that reflects the
interplay between many different mechanisms. It involves
weakness of the cardiac muscle that can be brought about
by genetic abnormalities, loss of cardiac muscle following
infarction, or secondary to pressure overload due to hyper-
tension. These stressful stimuli cause remodeling of the
heart which is initially compensatory but eventually leads to
ventricular dilatation, reduced cardiac output, and arrhyth-
mias. One of the features of cardiac failure is the presence of
high circulating levels of neurohumoral agents such as cat-
echolamines, angiotensin, endothelin, cytokines, insulin-
like growth factors, and natriuretic peptides (465). Overac-
tivity of the sympathetic nervous system and the renin-an-
giotensin-aldosterone system are particularly important,
and antagonists acting on these systems have proven valu-
able treatments that substantially reduce morbidity and
mortality (296). However, heart failure is characteristically
a chronic disease that is progressive so that even with recent
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
463Physiol Rev VOL 93 JANUARY 2013 www.prv.org
improvements in treatment novel approaches are required
to produce better outcomes.
Overactivity of the sympathetic nervous system is initially a
compensatory response that improves cardiac output by
increasing the force and rate of the heartbeat. Chronic ac-
tivation, however, leads to desensitization of
-adrenocep-
tors, cardiac hypertrophy, and apoptosis of cardiac myo-
cytes with fibrosis and cardiac arrhythmias. High levels of
angiotensin II and endothelin are associated with cardiac
hypertrophy, fibrosis, and cardiac arrhythmias. Blockade of
-adrenoceptors, ACE inhibitors, or AT
1
receptor blockade
are highly effective means of treating cardiac failure (296),
whereas the effects of endothelin antagonists are more
equivocal (207). Relaxin is now in phase III clinical trial for
the treatment of cardiac failure, yet the mechanism of action
is not well understood. Although there are a number of
possible mechanisms suggested by mainly animal studies,
clear links between these effects and the response to relaxin
in cardiac failure have yet to be made. The vasodilator
actions of relaxin mediated through NO (31, 33, 154, 343,
346) could have antihypertrophic effects (465), reduce pe-
ripheral resistance and hence cardiac work load, and also
help maintain viability of cardiac muscle. Relaxin is also
angiogenic, since cAMP increases in response to activation
of RXFP1 lead to increased expression of VEGF (541).
Relaxin has long been recognized as a powerful inotropic
and chronotropic agent (271) and has recently been shown
to have similar effects in humans (127). However, in all
species so far studied, these cardiac responses have been
confined to the atria and would be expected to have little
influence on ventricular ejection fraction. Relaxin also has
anti-inflammatory effects (345) and would be expected to
reduce tissue damage in cardiac failure. At present, it is not
clear which if any of these effects is involved in the positive
response to relaxin in cardiac failure. In addition, aspects of
relaxin-RXFP1 signaling have been extensively studied in in
vitro systems, but the situation could be quite different in path-
ological situations such as heart failure, particularly when fur-
ther complicated by additional drug administration. G pro-
tein-coupled receptor activation by catecholamines, angioten-
sin II, and endothelins leads to coupling to a variety of G
proteins including G
s
,G
q
,G
11
and G
␤␥
that in turn can
activate AC, phospholipase C, Akt, ERK1/2, p38MAPK, and
PKA to have hypertrophic effects (465). A compensatory re-
sponse that increases natriuretic peptides and leads to NO
stimulation of guanylyl cyclase causes an antihypertrophic
effect (465). Since relaxin is known to release NO and va-
sodilatation in systemic arteries is endothelium dependent
(170), this may underlie some of the beneficial effects of
relaxin. However, this study also showed that the effects of
relaxin were influenced by treatment of patients with ACE
inhibitors and responses of blood vessels to relaxin could be
altered by treatment with indomethacin or blockade of NO
synthase. In the phase I trial of relaxin, all of the patients
were treated with an ACE inhibitor or AT
1
receptor antag-
onist, as well as a diuretic and a
-adrenoceptor antagonist
(134). In the phase II clinical trial, 62–69% of patients were
treated with ACE inhibitors, 51–59% with
-adrenoceptor
antagonists, 24–38% with aldosterone antagonists, and
14–30% with digoxin (526). Clearly, there are many op-
portunities for interactions between relaxin and a variety of
pathways, and examination of potential interactions in in
vitro systems may help clarify the mechanism of action of
relaxin in vivo.
2. Is there a future for relaxin in the treatment
of fibrosis?
Although the early clinical trials of relaxin for the treatment
of sclerosis appeared encouraging (92) and a subsequent
placebo-controlled trial indicated that low-dose relaxin
produced significant improvement in skin and lung param-
eters (462), the phase II/III trial that followed was unsuc-
cessful (149, 461). It is possible that the failure of this trial
reflects the severity of the scleroderma in the patients chosen
for the trial (54). Nonetheless, the large number of animal
studies in which relaxin was successfully shown to have
antifibrotic properties still provide a strong case for its po-
tential utility in the treatment of pulmonary, renal, cardiac,
and hepatic fibrosis. Certainly work continues in this area
to attempt to demonstrate antifibrotic effects that will be
useful in humans (420, 437, 518).
D. Agonist and Antagonist Development for
RXFP Receptors
There is now a basic understanding of the mode of endog-
enous peptide ligand binding at RXFP1–3 with limited
knowledge of the mode of interaction of INSL5 with
RXFP4 (sect. III). There are currently peptide-based antag-
onists for all of the receptors with the most efficacious com-
pounds targeting RXFP3 (213, 299, 463). These com-
pounds have already been used to assist in determining the
role of relaxin-3 in the brain. Relaxin-3 peptide analogs
have also been developed that are specific agonists for
RXFP3 over RXFP1 (312, 463). However, these peptide-
based analogs are unlikely to cross the blood-brain barrier
so must be injected directly into the brain to be effective.
Currently there are no small molecule RXFP3 agonists or
antagonists reported, although a recent study has described
an RXFP3-specific relaxin-3-positive allosteric modulator
that was effective when used with relaxin-3 COOH-termi-
nal amide peptides but was unfortunately not effective with
the native relaxin COOH-terminal acid peptide (7).
The complex mechanism whereby relaxin binds to RXFP1 has
meant that there are currently no high-affinity peptide antag-
onists, and since the relaxin peptide itself rapidly loses activity
when truncated, it is likely that a small peptide agonist acting
at the orthosteric sites utilized by relaxin will be difficult to
develop (232, 233). However, what is clear is that the peptide
BATHGATE ET AL.
464 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
can be modified at the NH
2
terminus of both the A- and
B-chains, meaning that potentially it can be modified to im-
prove its pharmacokinetic properties and make it more drug-
like (232, 233). Small peptide antagonists have been devel-
oped that target RXFP2 that exploit the different binding
mechanism of INSL3 to RXFP2 compared with relaxin bind-
ing to RXFP1 (121, 464). However, the complex activation
mechanism associated with leucine-rich repeat contain-
ing GPCRs has precluded the development of small peptide
agonists. The isolated LDLa module has been demonstrated to
act as a RXFP1 antagonist (456, 457), and it is possible that if
the mechanism of action of the LDLa module in activating
RXFP1 and RXFP2 is determined, then specific agonists or
antagonists could be designed. Currently, no small molecule
RXFP1 or RXFP2 agonists or antagonists have been reported
(see NOTE ADDED IN PROOF 2). A possible approach that has
currently only been adopted for RXFP3 is the use of allosteric
modulators that typically target binding sites distinct from the
orthosteric sites utilized by the endogenous ligands.
An interesting recent development has been the discovery of
short linear peptides derived from an unidentified natural pre-
cursor protein that were able to activate RXFP1 and RXFP2,
but not RXFP3 receptors expressed in CHO cells (466, 467).
Further in vitro studies demonstrated that one of these pep-
tides (CGEN-25009) is able to activate fibroblasts and THP1
cells expressing RXFP1. The peptide was able to reduce lung
inflammation and injury as well as ameliorate adverse airway
remodeling and peribronchial fibrosis in a mouse model of
pulmonary fibrosis in a manner similar to relaxin (420).
CGEN-25009 and other peptides (CGEN-25010 and -25011)
produced complex concentration-response relationships in
cell systems expressing RXFP1 or RXFP2 with either inhibi-
tion or stimulation of cAMP accumulation being observed
(466). Some of the differences are attributable to assay condi-
tions such as the use of forskolin to amplify responses but also
could be explained by measurement of cAMP directly or by
the use of CRE reporter gene assays that display responses not
only to cAMP but also other signaling pathways and in par-
ticular MAPKs. It has also yet to be definitely demonstrated
that these peptides mediate their actions at RXFP1 or RXFP2
and even if they do the mechanism of binding is likely to be
very different from that of the endogenous ligands. However,
it is clear that some promising tools for the study of relaxin
family peptides are starting to emerge that will begin to answer
some of the questions raised in this section.
NOTE ADDED IN PROOF
1) A recent study has demonstrated that IN5L5/mice
are glucose intolerant and show elevated blood glucose lev-
els with advancing age. This results from a reduced area of
pancreatic islets and decreased numbers of
cells (Bur-
nicka-Turek et al., Endocrinology 153: 4655–4665, 2012).
2) Evidence was presented at the Relaxin 2012 meeting for
ML290, a small molecule agonist at RXFP1 which acts at a
site distinct from those at which relaxin binds (Chen et al. J
Biomol Screen doi:10.1177/1087057112469406.2012).
ACKNOWLEDGMENTS
We thank Dr. Bronwyn Evans for the analysis illustrated in
FIGURES 1 AND 2, Dr. Daniel Scott for FIGURE 2D, Dr.
Johan Rosengren for FIGURE 1C, Prof. Andrew Gundlach
and Dr. Craig Smith for contributions to FIGURE 8, and Dr.
Chrishan Samuel for helpful discussions. Some of the tissue
illustrations in FIGURE 4 were downloaded from www.med.
ars.it with permission from Dr. Davide Brunelli.
R. A. D. Bathgate and M. L. Halls contributed equally to
the writing of this review.
Present addresses: R. A. D. Bathgate, M. L. Halls, M. Ko-
can, and R. J. Summers, Drug Discovery Biology, Monash
Institute of Pharmaceutical Sciences & Department of Phar-
macology, Monash University, Victoria, Australia; G. E.
Callander, Florey Neurosciences Institutes and Department
of Biochemistry and Molecular Biology, The University of
Melbourne, Victoria, Australia; and E. T. van der Westhui-
zen, Université de Montréal, Institut de Recherche en Im-
munologie et Cancérologie, Pharmacologie Moleculaire,
Montréal, Canada.
Address for reprint requests and other correspondence: R. J.
Summers, Drug Discovery Biology, Monash Institute of
Pharmaceutical Sciences & Department of Pharmacology,
Monash University, 399 Royal Parade, Victoria 3052, Aus-
tralia (e-mail: roger.summers@monash.edu).
GRANTS
This work was funded by the National Health and Medical
Research Council (NHMRC) of Australia Program Grant
519461 (to P. M. Sexton, A. Christopoulos, and R. J. Sum-
mers), NHMRC Career Development Awards 519581 (to
M. L. Halls) and 1013819 (to E. Van Der Westhuizen),
NHMRC Senior Research Fellowship 509011 (to R. A. D.
Bathgate), and NHMRC Project Grant 628427 (to R. A. D.
Bathgate).
DISCLOSURES
No conflicts of interest, financial or otherwise, are declared
by the authors.
REFERENCES
1. Adams WC, Frieden EH. Inhibition of postpartum uterine involution in the rat by
relaxin. Biol Reprod 33: 1168 –1175, 1985.
2. Adham IM, Burkhardt E, Benahmed M, Engel W. Cloning of a cDNA for a novel
insulin-like peptide of the testicular Leydig cells. J Biol Chem 268: 26668 –26672, 1993.
3. Adham IM, Steding G, Thamm T, Bullesbach EE, Schwabe C, Paprotta I, Engel W. The
overexpression of the INSL3 in female mice causes descent of the ovaries. Mol Endo-
crinol 16: 244 –252, 2002.
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
465Physiol Rev VOL 93 JANUARY 2013 www.prv.org
4. Ahmad N, Wang W, Nair R, Kapila S. Relaxin induces matrix-metalloproteinases-9 and
-13 via RXFP1: induction of MMP-9 involves the PI3K, ERK, Akt and PKC-zeta path-
ways. Mol Cell Endocrinol 363: 46 – 61, 2012.
5. Alexander BT, Miller MT, Kassab S, Novak J, Reckelhoff JF, Kruckeberg WC, Granger
JP. Differential expression of renal nitric oxide synthase isoforms during pregnancy in
rats. Hypertension 33: 435– 439, 1999.
6. Alleva DG, Kaser SB, Monroy MA, Fenton MJ, Beller DI. IL-15 functions as a potent
autocrine regulator of macrophage proinflammatory cytokine production: evidence
for differential receptor subunit utilization associated with stimulation or inhibition. J
Immunol 159: 2941–2951, 1997.
7. Alvarez-Jaimes L, Sutton SW, Nepomuceno D, Motley ST, Cik M, Stocking E, Shob-
lock J, Bonaventure P. In vitro pharmacological characterization of RXFP3 allosterism:
an example of probe dependency. PloS one 7: e30792, 2012.
8. Amory JK, Page ST, Anawalt BD, Coviello AD, Matsumoto AM, Bremner WJ. Elevated
end-of-treatment serum INSL3 is associated with failure to completely suppress sper-
matogenesis in men receiving male hormonal contraception. J Androl 28: 548 –554,
2007.
9. Anand-Ivell R, Heng K, Hafen B, Setchell B, Ivell R. Dynamics of INSL3 peptide
expression in the rodent testis. Biol Reprod 81: 480 – 487, 2009.
10. Anand-Ivell R, Ivell R, Driscoll D, Manson J. Insulin-like factor 3 levels in amniotic fluid
of human male fetuses. Hum Reprod 23: 1180 –1186, 2008.
11. Anand-Ivell R, Wohlgemuth J, Haren MT, Hope PJ, Hatzinikolas G, Wittert G, Ivell R.
Peripheral INSL3 concentrations decline with age in a large population of Australian
men. Int J Androl 29: 618 – 626, 2006.
12. Anand-Ivell RJ, Relan V, Balvers M, Coiffec-Dorval I, Fritsch M, Bathgate RA, Ivell R.
Expression of the insulin-like peptide 3 (INSL3) hormone-receptor (LGR8) system in
the testis. Biol Reprod 74: 945–953, 2006.
13. Anand-Ivell RJK, Heng K, Bartsch O, Ivell R. Relaxin signaling in THP-1 cells uses a
novel phosphotyrosine-dependent pathway. Mol Cell Endocrinol 272: 1–13, 2007.
14. Anderson LL, Bast JD, Melampy RM. Relaxin in ovarian tissue during different repro-
ductive stages in the rat. J Endocrinol 59: 371–372, 1973.
15. Anderson LL, Ford JJ, Melampy RM, Cox DF. Relaxin in porcine corpora lutea during
pregnancy and after hysterectomy. Am J Physiol 225: 1215–1219, 1973.
16. Anwer K, Hovington JA, Sanborn BM. Involvement of protein kinase A in the regula-
tion of intracellular free calcium and phosphoinositide turnover in rat myometrium.
Biol Reprod 43: 851– 859, 1990.
17. Arora A, Scholar EM. Role of tyrosine kinase inhibitors in cancer therapy. J Pharmacol
Exp Ther 315: 971–979, 2005.
18. Ascoli M, Fanelli F, Segaloff DL. The lutropin/choriogonadotropin receptor, a 2002
perspective. Endocr Rev 23: 141–174, 2002.
19. Baccari MC, Bani D, Bigazzi M, Calamai F. Influence of relaxin on the neurally induced
relaxant responses of the mouse gastric fundus. Biol Reprod 71: 1325–1329, 2004.
20. Baccari MC, Nistri S, Vannucchi MG, Calamai F, Bani D. Reversal by relaxin of altered
ileal spontaneous contractions in dystrophic (mdx) mice through a nitric oxide-medi-
ated mechanism. Am J Physiol Regul Integr Comp Physiol 293: R662–R668, 2007.
21. Baillie GS, Scott JD, Houslay MD. Compartmentalization of phosphodiesterases and
protein kinase A: opposites attract. FEBS Lett 579: 3264 –3270, 2005.
22. Baker JG, Hill SJ. Multiple GPCR conformations and signaling pathways: implications
for antagonist affinity estimates. Trends Pharmacol Sci 28: 374 –381, 2007.
23. Balvers M, Spiess AN, Domagalski R, Hunt N, Kilic E, Mukhopadhyay AK, Hanks E,
Charlton HM, Ivell R. Relaxin-like factor expression as a marker of differentiation in
the mouse testis and ovary. Endocrinology 136: 2960 –2970, 1998.
24. Bamberger AM, Ivell R, Balvers M, Kelp B, Bamberger CM, Riethdorf L, Loning T.
Relaxin-like factor (RLF): a new specific marker for Leydig cells in the ovary. Int J
Gynecol Pathol 18: 163–168, 1999.
25. Banerjee A, Shen PJ, Ma S, Bathgate RA, Gundlach AL. Swim stress excitation of
nucleus incertus and rapid induction of relaxin-3 expression via CRF1 activation.
Neuropharmacology 58: 145–155, 2010.
26. Bani D. Relaxin and breast cancer. Bull Cancer 84: 179 –182, 1997.
27. Bani D, Baccari MC, Nistri S, Calamai F, Bigazzi M, Sacchi TB. Relaxin up-regulates the
nitric oxide biosynthetic pathway in the mouse uterus: involvement in the inhibition of
myometrial contractility. Endocrinology 140: 4434 – 4441, 1999.
28. Bani D, Baccari MC, Quattrone S, Nistri S, Calamai F, Bigazzi M, Bani Sacchi T. Relaxin
depresses small bowel motility through a nitric oxide-mediated mechanism: studies in
mice. Biol Reprod 66: 778 –784, 2002.
29. Bani D, Baronti R, Vannacci A, Bigazzi M, Sacchi TB, Mannaioni PF, Masini E. Inhibitory
effects of relaxin on human basophils activated by stimulation of the Fc epsilon recep-
tor. The role of nitric oxide. Int Immunopharmacol 2: 1195–1204, 2002.
30. Bani D, Bigazzi M, Masini E, Bani G, Sacchi TB. Relaxin depresses platelet aggre-
gation: in vitro studies on isolated human and rabbit platelets. Lab Invest 73:
709 –716, 1995.
31. Bani D, Failli P, Bello MG, Thiemermann C, Bani Sacchi T, Bigazzi M, Masini E. Relaxin
activates the L-arginine-nitric oxide pathway in vascular smooth muscle cells in cul-
ture. Hypertension 31: 1240 –1247, 1998.
32. Bani D, Masini E, Bello MG, Bigazzi M, Sacchi TB. Relaxin activates the L-arginine-nitric
oxide pathway in human breast cancer cells. Cancer Res 55: 5272–5275, 1995.
33. Bani D, Masini E, Bello MG, Bigazzi M, Sacchi TB. Relaxin protects against myocardial
injury caused by ischemia and reperfusion in rat heart. Am J Pathol 152: 1367–1376,
1998.
34. Bani D, Nistri S, Quattrone S, Bigazzi M, Sacchi TB. Relaxin causes changes of the liver.
In vivo studies in rats. Horm Metab Res 33: 175–180, 2001.
35. Bani G, Bani ST, Bigazzi M, Bianchi S. Effects of relaxin on the microvasculature of
mouse mammary gland. Histol Histopathol 3: 337–343, 1988.
36. Bani-Sacchi T, Bigazzi M, Bani D, Mannaioni PF, Masini E. Relaxin-induced increased
coronary flow through stimulation of nitric oxide production. Br J Pharmacol 116:
1589 –1594, 1995.
37. Bartsch O, Bartlick B, Ivell R. Phosphodiesterase 4 inhibition synergizes with relaxin
signaling to promote decidualization of human endometrial stromal cells. J Clin Endo-
crinol Metab 89: 324 –334, 2004.
38. Bartsch O, Bartlick B, Ivell R. Relaxin signaling links tyrosine phosphorylation to phos-
phodiesterase and adenylyl cyclase activity. Mol Hum Reprod 7: 799 – 809, 2001.
39. Bates P, Young JAT, Varmus HE. A receptor for subgroup A Rous sarcoma virus is
related to the low density lipoprotein receptor. Cell 74: 1043–1051, 1993.
40. Bathgate R, Balvers M, Hunt N, Ivell R. Relaxin-like factor gene is highly expressed in
the bovine ovary of the cycle and pregnancy: sequence and messenger ribonucleic
acid analysis. Biol Reprod 55: 1452–1457, 1996.
41. Bathgate R, Moniac N, Bartlick B, Schumacher M, Fields M, Ivell R. Expression and
regulation of relaxin-like factor gene transcripts in the bovine ovary: differentiation-
dependent expression in theca cell cultures. Biol Reprod 61: 1090 –1098, 1999.
42. Bathgate RA, Ivell R, Sanborn BM, Sherwood OD, Summers RJ. International Union of
Pharmacology LVII: Recommendations for the nomenclature of receptors for relaxin
family peptides. Pharmacol Rev 58: 7–31, 2006.
43. Bathgate RA, Ivell R, Sanborn BM, Sherwood OD, Summers RJ. Receptors for relaxin
family peptides. Ann NY Acad Sci 1041: 61–76, 2005.
44. Bathgate RA, Lekgabe ED, McGuane JT, Su Y, Pham T, Ferraro T, Layfield S, Hannan
RD, Thomas WG, Samuel CS, Du XJ. Adenovirus-mediated delivery of relaxin re-
verses cardiac fibrosis. Mol Cell Endocrinol 280: 30 –38, 2008.
45. Bathgate RA, Lin F, Hanson NF, Otvos L Jr, Guidolin A, Giannakis C, Bastiras S,
Layfield SL, Ferraro T, Ma S, Zhao C, Gundlach AL, Samuel CS, Tregear GW, Wade
JD. Relaxin-3: improved synthesis strategy and demonstration of its high-affinity in-
teraction with the relaxin receptor LGR7 both in vitro and in vivo. Biochemistry 45:
1043–1053, 2006.
46. Bathgate RA, Samuel CS, Burazin TC, Layfield S, Claasz AA, Reytomas IG, Dawson
NF, Zhao C, Bond CP, Summers RJ, Parry LJ, Wade JD, Tregear GW. Human relaxin
gene 3 (H3) and the equivalent mouse relaxin (M3) gene. Novel members of the
relaxin peptide family. J Biol Chem 277: 1148 –1157, 2002.
BATHGATE ET AL.
466 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
47. Bathgate RAD, Hsueh AJW, Sherwood OD. Physiology and molecular biology of the
relaxin peptide family. In: Knobil and Neill’s Physiology of Reproduction , edited by Neill
JD. New York: Academic, 2006.
48. Bathgate RAD, Moniac N, Bartlick B, Balvers M, Domagalski R, Dawson N, Wade JD,
Tregear GW, Ivell R. The relaxin-like factor (RLF, Insl3) is highly expressed in the
ruminant ovary. In: Relaxin 2000: Proceedings of the Third International Conference on
Relaxin and Related Peptides , edited by Tregear GW, Ivell R, Bathgate RA, Wade JD.
Dordrecht, Germany: Kluwer, 2001, p. 349 –356.
49. Bathgate RAD, Scott D, Chung S, Ellyard D, Garreffa A, Tregear GW. Searching the
human genome database for novel relaxin-like peptides. Letts Pept Sci 8: 129 –132,
2002.
50. Bay K, Hartung S, Ivell R, Schumacher M, Jurgensen D, Jorgensen N, Holm M, Skak-
kebaek NE, Andersson AM. Insulin-like factor 3 serum levels in 135 normal men and
85 men with testicular disorders: relationship to the luteinizing hormone-testoster-
one axis. J Clin Endocrinol Metab 90: 3410 –3418, 2005.
51. Bell RJ, Eddie LW, Lester AR, Wood EC, Johnston PD, Niall HD. Relaxin in human
pregnancy serum measured with an homologous radioimmunoassay. Obstet Gynecol
69: 585–589, 1987.
52. Bell RJ, Permezel M, MacLennan A, Hughes C, Healy D, Brennecke S. A randomized,
double-blind, placebo-controlled trial of the safety of vaginal recombinant human
relaxin for cervical ripening. Obstet Gynecol 82: 328 –333, 1993.
53. Belt WD, Anderson LL, Cavazos LF, Melampy RM. Cytoplasmic granules and relaxin
levels in porcine corpora lutea. Endocrinology 89: 1–10, 1971.
54. Bennett RG. Relaxin and its role in the development and treatment of fibrosis. Transl
Res 154: 1– 6, 2009.
55. Bennett RG, Kharbanda KK, Tuma DJ. Inhibition of markers of hepatic stellate cell
activation by the hormone relaxin. Biochem Pharmacol 66: 867– 874, 2003.
56. Bieche I, Laurent A, Laurendeau I, Duret L, Giovangrandi Y, Frendo JL, Olivi M,
Fausser JL, Evain-Brion D, Vidaud M. Placenta-specific INSL4 expression is mediated
by a human endogenous retrovirus element. Biol Reprod 68: 1422–1429, 2003.
57. Bigazzi M, Brandi ML, Bani G, Sacchi TB. Relaxin influences the growth of MCF-7
breast cancer cells. Mitogenic and antimitogenic action depends on peptide concen-
tration. Cancer 70: 639 – 643, 1992.
58. Bigazzi M, Del MA, Petrucci F, Casali R, Novelli GP. The local administration of relaxin
induces changes in the microcirculation of the rat mesocaecum. Acta Endocrinol Co-
penh 112: 296 –299, 1986.
59. Binder C, Hagemann T, Husen B, Schulz M, Einspanier A. Relaxin enhances in-vitro
invasiveness of breast cancer cell lines by up-regulation of matrix metalloproteases.
Mol Hum Reprod 8: 789 –796, 2002.
60. Binder C, Simon A, Binder L, Hagemann T, Schulz M, Emons G, Trumper L, Eins-
panier A. Elevated concentrations of serum relaxin are associated with metastatic
disease in breast cancer patients. Breast Cancer Res Treat 87: 157–166, 2004.
61. Blankenship T, Stewart DR, Benirschke K, King B, Lasley BL. Immunocytochemical
localization of nonluteal ovarian relaxin. J Reprod Med 39: 235–240, 1994.
62. Boels K, Hermans-Borgmeyer I, Schaller HC. Identification of a mouse orthologue of
the G-protein-coupled receptor SALPR and its expression in adult mouse brain and
during development. Brain Res Dev Brain Res 152: 265–268, 2004.
63. Boels K, Schaller HC. Identification and characterization of GPR100 as a novel human
G-protein-coupled bradykinin receptor. Br J Pharmacol 140: 932–938, 2003.
64. Bogatcheva NV, Ferlin A, Feng S, Truong A, Gianesello L, Foresta C, Agoulnik AI.
T222P mutation of the insulin-like 3 hormone receptor LGR8 is associated with
testicular maldescent and hinders receptor expression on the cell surface membrane.
Am J Physiol Endocrinol Metab 292: E138 –E144, 2007.
65. Bogatcheva NV, Truong A, Feng S, Engel W, Adham IM, Agoulnik AI. GREAT/LGR8 is
the only receptor for insulin-like 3 peptide. Mol Endocrinol 17: 2639 –2646, 2003.
66. Bohm M, Eschenhagen T, Gierschik P, Larisch K, Lensche H, Mende U, Schmitz W,
Schnabel P, Scholz H, Steinfath M. Radioimmunochemical quantification of Gi alpha in
right and left ventricles from patients with ischaemic and dilated cardiomyopathy and
predominant left ventricular failure. J Mol Cell Cardiol 26: 133–149, 1994.
67. Bond CP, Parry LJ, Samuel CS, Gehring HM, Lederman FL, Rogers PA, Summers RJ.
Increased expression of the relaxin receptor (LGR7) in human endometrium during
the secretory phase of the menstrual cycle. J Clin Endocrinol Metab 89: 3477–3485,
2004.
68. Boockfor FR, Fullbright G, Bullesbach EE, Schwabe C. Relaxin-like factor (RLF) serum
concentrations and gubernaculum RLF receptor display in relation to pre- and neo-
natal development of rats. Reproduction 122: 899 –906, 2001.
69. Braddon SA. Relaxin-dependent adenosine 6=,5=-monophosphate concentration
changes in the mouse pubic symphysis. Endocrinology 102: 1292–1299, 1978.
70. Bryant GG, Rees MC, Turnbull AC. Immunohistochemical localization of relaxin,
prolactin and prostaglandin synthase in human amnion, chorion and decidua. J Endo-
crinol 114: 491– 496, 1987.
71. Bryant-Greenwood GD, Rutanen EM, Partanen S, Coelho TK, Yamamoto SY. Se-
quential appearance of relaxin, prolactin and IGFBP-1 during growth and differentia-
tion of the human endometrium. Mol Cell Endocrinol 95: 23–29, 1993.
72. Büllesbach EE, Rhodes R, Rembiesa B, Schwabe C. The relaxin-like factor is a hor-
mone. Endocrine 10: 167–169, 1999.
73. Bullesbach EE, Schwabe C. Functional importance of the A chain loop in relaxin and
insulin. J Biol Chem 269: 13124 –13128, 1994.
74. Bullesbach EE, Schwabe C. LGR8 signal activation by the relaxin-like factor. J Biol
Chem 280: 14856 –14890, 2005.
75. Bullesbach EE, Schwabe C. The mode of interaction of the relaxin-like factor (RLF)
with the leucine-rich repeat G protein-activated receptor 8. J Biol Chem 281: 26136 –
26143, 2006.
76. Bullesbach EE, Schwabe C. The primary structure and the disulfide links of the bovine
relaxin-like factor (RLF). Biochemistry 41: 274 –281, 2002.
77. Bullesbach EE, Schwabe C. Synthetic cross-links arrest the COOH-terminal region of
the relaxin-like factor in an active conformation. Biochemistry 43: 8021– 8028, 2004.
78. Bullesbach EE, Schwabe C. The trap-like relaxin-binding site of LGR7. J Biol Chem 280:
14051–14056, 2005.
79. Büllesbach EE, Schwabe C. Synthesis and conformational analysis of the insulin-like 4
gene product. J Pept Res 57: 77– 83, 2001.
80. Burazin TC, Bathgate RA, Macris M, Layfield S, Gundlach AL, Tregear GW. Restricted,
but abundant, expression of the novel rat gene-3 (R3) relaxin in the dorsal tegmental
region of brain. J Neurochem 82: 1553–1557, 2002.
81. Burazin TC, Johnson KJ, Ma S, Bathgate RA, Tregear GW, Gundlach AL. Localization
of LGR7 (relaxin receptor) mRNA and protein in rat forebrain: correlation with
relaxin binding site distribution. Ann NY Acad Sci 1041: 205–210, 2005.
82. Burger LL, Sherwood OD. Evidence that cellular proliferation contributes to relaxin-
induced growth of both the vagina and the cervix in the pregnant rat. Endocrinology
136: 4820 – 4826, 1995.
83. Burger LL, Sherwood OD. Relaxin increases the accumulation of new epithelial and
stromal cells in the rat cervix during the second half of pregnancy. Endocrinology 139:
3984 –3995, 1998.
84. Burkhardt E, Adham IM, Hobohm U, Murphy D, Sander C, Engel W. A human cDNA
coding for the Leydig insulin-like peptide (Ley I-L). Hum Genet 94: 91–94, 1994.
85. Burnicka-Turek O, Shirneshan K, Paprotta I, Grzmil P, Meinhardt A, Engel W, Adham
IM. Inactivation of insulin-like factor 6 disrupts the progression of spermatogenesis at
late meiotic prophase. Endocrinology 150: 4348 – 4357, 2009.
86. Callander GE, Bathgate RA. Relaxin family peptide systems and the central nervous
system. Cell Mol Life Sci 67: 2327–2341, 2010.
87. Callander GE, Thomas WG, Bathgate RA. Prolonged RXFP1 and RXFP2 signaling can
be explained by poor internalization and a lack of beta-arrestin recruitment. Am J
Physiol Cell Physiol 296: C1058 –C1066, 2009.
88. Campitiello MR, De Franciscis P, Mele D, Izzo G, Sinisi A, Delrio G, Colacurci N.
Endometrial LGR7 expression during menstrual cycle. Fertil Steril 95: 2511–2514,
2011.
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
467Physiol Rev VOL 93 JANUARY 2013 www.prv.org
89. Cardoso LC, Nascimento AR, Royer C, Porto CS, Lazari MF. Locally produced relaxin
may affect testis and vas deferens function in rats. Reproduction 139: 185–196, 2010.
90. Carmon KS, Gong X, Lin Q, Thomas A, Liu Q. R-spondins function as ligands of the
orphan receptors LGR4 and LGR5 to regulate Wnt/{beta}-catenin signaling. Proc Natl
Acad Sci USA 2011.
91. Carrell DT, Peterson CM, Urry RL. The binding of recombinant human relaxin to
human spermatozoa. Endocrinol Res 21: 697–707, 1995.
92. Casten GG, Boucek RJ. Use of relaxin in the treatment of scleroderma. J Am Med Assoc
166: 319 –324, 1958.
93. Chabre M, Deterre P, Antonny B. The apparent cooperativity of some GPCRs does
not necessarily imply dimerization. Trends Pharmacol Sci 30: 182–187, 2009.
94. Chassin D, Laurent A, Janneau JL, Berger R, Bellet D. Cloning of a new member of the
insulin gene superfamily (INSL4) expressed in human placenta. Genomics 29: 465–
470, 1995.
95. Cheah SH, Sherwood OD. Target tissues for relaxin in the rat: tissue distribution of
injected 125I-labeled relaxin and tissue changes in adenosine 3=,5=-monophosphate
levels after in vitro relaxin incubation. Endocrinology 106: 1203–1209, 1980.
96. Chen GA, Huang JR, Tseng L. The effect of relaxin on cyclic adenosine 3=,5=-mono-
phosphate concentrations in human endometrial glandular epithelial cells. Biol Reprod
39: 519 –525, 1988.
97. Chen GR, Dong L, Ge RS, Hardy MP. Relationship between phthalates and testicular
dysgenesis syndrome. Zhonghua Nan Ke Xue 13: 195–200, 2007.
98. Chen J, Kuei C, Sutton SW, Bonaventure P, Nepomuceno D, Eriste E, Sillard R,
Lovenberg TW, Liu C. Pharmacological characterization of relaxin-3/INSL7 receptors
GPCR135 and GPCR142 from different mammalian species. J Pharmacol Exp Ther 312:
83–95, 2005.
99. Chen J, Qiu Q, Lohstroh PN, Overstreet JW, Lasley BL. Hormonal characteristics in
the early luteal phase of conceptive and nonconceptive menstrual cycles. J Soc Gynecol
Invest 10: 27–31, 2003.
100. Conklin D, Lofton-Day CE, Haldeman BA, Ching A, Whitmore TE, Lok S, Jaspers S.
Identification of INSL5, a new member of the insulin superfamily. Genomics 60: 50 –
56, 1999.
101. Conrad KP. Maternal vasodilation in pregnancy: the emerging role of relaxin. Am J
Physiol Regul Integr Comp Physiol 301: R267–R275, 2011.
102. Conrad KP. Unveiling the vasodilatory actions and mechanisms of relaxin. Hyperten-
sion 56: 2–9, 2010.
103. Conrad KP, Debrah DO, Novak J, Danielson LA, Shroff SG. Relaxin modifies systemic
arterial resistance and compliance in conscious, nonpregnant rats. Endocrinology 145:
3289 –3296, 2004.
104. Conrad KP, Novak J. Emerging role of relaxin in renal and cardiovascular function. Am
J Physiol Regul Integr Comp Physiol 287: R250 –R261, 2004.
105. Conrad KP, Shroff SG. Effects of relaxin on arterial dilation, remodeling, and mechan-
ical properties. Curr Hypertens Reports 13: 409 – 420, 2011.
106. Cools R, Roberts AC, Robbins TW. Serotoninergic regulation of emotional and be-
havioural control processes. Trends Cogn Sci 12: 31– 40, 2008.
107. Coulson CC, Thorp JM Jr, Mayer DC, Cefalo RC. Central hemodynamic effects of
recombinant human relaxin in the isolated, perfused rat heart model. Obstet Gynecol
87: 610 – 612, 1996.
108. Crish JF, Soloff MS, Shaw AR. Changes in relaxin precursor mRNA levels in the rat
ovary during pregnancy. J Biol Chem 261: 1909 –1913, 1986.
109. Cronin MJ, Malaska T, Bakhit C. Human relaxin increases cyclic AMP levels in cultured
anterior pituitary cells. Biochem Biophys Res Commun 148: 1246 –1251, 1987.
110. Daly NL, Scanlon MJ, Djordjevic JT, Kroon PA, Smith R. Three-dimensional structure
of a cysteine-rich repeat from the low-density lipoprotein receptor. Prac Natl Acad Sci
USA 92: 6334 – 6338, 1995.
111. Danielson LA, Conrad KP. Acute blockade of nitric oxide synthase inhibits renal
vasodilation and hyperfiltration during pregnancy in chronically instrumented con-
scious rats. J Clin Invest 96: 482– 490, 1995.
112. Danielson LA, Conrad KP. Time course and dose response of relaxin-mediated renal
vasodilation, hyperfiltration, and changes in plasma osmolality in conscious rats. J Appl
Physiol 95: 1509 –1514, 2003.
113. Danielson LA, Kercher LJ, Conrad KP. Impact of gender and endothelin on renal
vasodilation and hyperfiltration induced by relaxin in conscious rats. Am J Physiol Regul
Integr Comp Physiol 279: R1298 –R1304, 2000.
114. Danielson LA, Sherwood OD, Conrad KP. Relaxin is a potent renal vasodilator in
conscious rats. J Clin Invest 103: 525–533, 1999.
115. Davis D, Liu X, Segaloff DL. Identification of the sites of N-linked glycosylation on the
follicle-stimulating hormone (FSH) receptor and assessment of their role in FSH
receptor function. Mol Endocrinol 9: 159 –170, 1995.
116. Davis RJ. Signal transduction by the JNK group of MAP kinases. Cell 103: 239–252,
2000.
117. De Cooman S, Gilliaux P, Thomas K. Immunoreactive relaxin-like substance in human
split ejaculates. Fertil Steril 39: 111–113, 1983.
118. Debrah DO, Conrad KP, Danielson LA, Shroff SG. Effects of relaxin on systemic
arterial hemodynamics and mechanical properties in conscious rats: sex dependency
and dose response. J Appl Physiol 98: 1013–1020, 2005.
119. Debrah DO, Novak J, Matthews JE, Ramirez RJ, Shroff SG, Conrad KP. Relaxin is
essential for systemic vasodilation and increased global arterial compliance during
early pregnancy in conscious rats. Endocrinology 147: 5126 –5131, 2006.
120. DeFea KA. Beta-arrestins as regulators of signal termination and transduction: how do
they determine what to scaffold? Cell Signal 23: 621– 629, 2011.
121. Del Borgo MP, Hughes RA, Bathgate RA, Lin F, Kawamura K, Wade JD. Analogs of
insulin-like peptide 3 (INSL3) B-chain are LGR8 antagonists in vitro and in vivo. J Biol
Chem 281: 13068 –13074, 2006.
122. Dessauer CW. Adenylyl cyclase-A-kinase anchoring protein complexes: the next di-
mension in cAMP signaling. Mol Pharmacol 76: 935–941, 2009.
123. DiScipio RG, Gehring MR, Podack ER, Kan CC, Hugli TE, Fey GH. Nucleotide se-
quence of cDNA and derived amino acid sequence of human complement component
C9. Prac Natl Acad Sci USA 81: 7298 –7302, 1984.
124. Dodge KL, Carr DW, Sanborn BM. Protein kinase A anchoring to the myometrial
plasma membrane is required for cyclic adenosine 3=,5=-monophosphate regulation
of phosphatidylinositide turnover. Endocrinology 140: 5165–5170, 1999.
125. Donizetti A, Grossi M, Pariante P, D’Aniello E, Izzo G, Minucci S, Aniello F. Two
neuron clusters in the stem of postembryonic zebrafish brain specifically express
relaxin-3 gene: first evidence of nucleus incertus in fish. Dev Dyn 237: 3864 –3869,
2008.
126. Drolet DW, Henzel WJ, Johnston PD. Purification, amino-terminal sequencing and
demonstration of biological activity of human relaxin from corpora lutea. In: 69th
Annual Meeting of the Endocrine Society. 1987, p. 196.
127. Dschietzig T, Alexiou K, Kinkel HT, Baumann G, Matschke K, Stangl K. The positive
inotropic effect of relaxin-2 in human atrial myocardium is preserved in end-stage
heart failure: role of G(i)-phosphoinositide-3 kinase signaling. J Card Fail 17: 158 –166,
2011.
128. Dschietzig T, Bartsch C, Baumann G, Stangl K. RXFP1-inactive relaxin activates hu-
man glucocorticoid receptor: further investigations into the relaxin-GR pathway.
Regul Pept 154: 77– 84, 2009.
129. Dschietzig T, Bartsch C, Richter C, Laule M, Baumann G, Stangl K. Relaxin, a preg-
nancy hormone, is a functional endothelin-1 antagonist: attenuation of endothelin-1-
mediated vasoconstriction by stimulation of endothelin type-B receptor expression
via ERK-1/2 and nuclear factor-kappaB. Circ Res 92: 32– 40, 2003.
130. Dschietzig T, Bartsch C, Stangl V, Baumann G, Stangl K. Identification of the preg-
nancy hormone relaxin as glucocorticoid receptor agonist. FASEB J 18: 1536–1538,
2004.
131. Dschietzig T, Bartsch C, Wessler S, Baumann G, Stangl K. Autoregulation of human
relaxin-2 gene expression critically involves relaxin and glucocorticoid receptor bind-
ing to glucocorticoid response half-sites in the relaxin-2 promoter. Regul Pept 155:
163–173, 2009.
BATHGATE ET AL.
468 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
132. Dschietzig T, Richter C, Bartsch C, Laule M, Armbruster FP, Baumann G, Stangl K.
The pregnancy hormone relaxin is a player in human heart failure. FASEB J 15: 2187–
2195, 2001.
133. Dschietzig T, Teichman S, Unemori E, Wood S, Boehmer J, Richter C, Baumann G,
Stangl K. First clinical experience with intravenous recombinant human relaxin in
compensated heart failure. Ann NY Acad Sci 1160: 387–392, 2009.
134. Dschietzig T, Teichman S, Unemori E, Wood S, Boehmer J, Richter C, Baumann G,
Stangl K. Intravenous recombinant human relaxin in compensated heart failure: a
safety, tolerability, and pharmacodynamic trial. J Card Fail 15: 182–190, 2009.
135. Du XJ, Bathgate RA, Samuel CS, Dart AM, Summers RJ. Cardiovascular effects of
relaxin: from basic science to clinical therapy. Nat Rev Cardiol 7: 48 –58, 2010.
136. Du XJ, Samuel CS, Gao XM, Zhao L, Parry LJ, Tregear GW. Increased myocardial
collagen and ventricular diastolic dysfunction in relaxin deficient mice: a gender-
specific phenotype. Cardiovasc Res 57: 395– 404, 2003.
137. Dun SL, Brailoiu E, Wang Y, Brailoiu GC, Liu-Chen LY, Yang J, Chang JK, Dun NJ.
Insulin-like peptide 5: expression in the mouse brain and mobilization of calcium.
Endocrinology 147: 3243–3248, 2006.
138. Eddie LW, Bell RJ, Lester A, Geier M, Bennett G, Johnston PD, Niall HD. Radioim-
munoassay of relaxin in pregnancy with an analogue of human relaxin. Lancet 1:
1344 –1346, 1986.
139. Eddie LW, Cameron IT, Leeton JF, Healy DL, Renou P. Ovarian relaxin is not essential
for dilatation of cervix. Lancet 336: 243, 1990.
140. Eddie LW, Martinez F, Healy DL, Sutton B, Bell RJ, Tregear GW. Relaxin in sera during
the luteal phase of in-vitro fertilization cycles. Br J Obstet Gynaecol 97: 215–220, 1990.
141. Eddie LW, Sutton B, Fitzgerald S, Bell RJ, Johnston PD, Tregear GW. Relaxin in paired
samples of serum and milk from women after term and preterm delivery. Am J Obstet
Gynecol 161: 970 –973, 1989.
142. Efendiev R, Samelson BK, Nguyen BT, Phatarpekar PV, Baameur F, Scott JD, Des-
sauer CW. AKAP79 interacts with multiple adenylyl cyclase (AC) isoforms and scaf-
folds AC5 and -6 to alpha-amino-3-hydroxyl-5-methyl-4-isoxazole-propionate
(AMPA) receptors. J Biol Chem 285: 14450 –14458, 2010.
143. Egan CT, Herrick-Davis K, Teitler M. Creation of a constitutively activated state of the
5-hydroxytryptamine
2A
receptor by site-directed mutagenesis: inverse agonist activ-
ity of antipsychotic drugs. J Pharmacol Exp Ther 286: 85–90, 1998.
144. Eigenbrot C, Randal M, Quan C, Burnier J, O’Connell L, Rinderknecht E, Kossiakoff
AA. X-ray structure of human relaxin at 1.5 A. Comparison to insulin and implications
for receptor binding determinants. J Mol Biol 221: 15–21, 1991.
145. Einspanier A, Muller D, Lubberstedt J, Bartsch O, Jurdzinski A, Fuhrmann K, Ivell R.
Characterization of relaxin binding in the uterus of the marmoset monkey. Mol Hum
Reprod 7: 963–970, 2001.
146. Einspanier A, Nubbemeyer R, Schlote S, Schumacher M, Ivell R, Fuhrmann K, Marten
A. Relaxin in the marmoset monkey: secretion pattern in the ovarian cycle and early
pregnancy. Biol Reprod 61: 512–520, 1999.
147. Emmen JM, McLuskey A, Adham IM, Engel W, Grootegoed JA, Brinkmann AO. Hor-
monal control of gubernaculum development during testis descent: gubernaculum
outgrowth in vitro requires both insulin-like factor and androgen. Endocrinology 141:
4720 – 4727, 2000.
148. Emmi AM, Skurnick J, Goldsmith LT, Gagliardi CL, Schmidt CL, Kleinberg D, Weiss G.
Ovarian control of pituitary hormone secretion in early human pregnancy. J Clin
Endocrinol Metab 72: 1359 –1363, 1991.
149. Erikson MS, Unemori EN. Relaxin clinical trials in systemic sclerosis. In: Relaxin 2000:
Proceedings of the Third International Conference on Relaxin and Related Peptides , edited
by Tregear GW, Ivell R, Bathgate RA, Wade JD. Amsterdam: Kluwer, 2001, p.
373–382.
150. Eschenhagen T, Mende U, Nose M, Schmitz W, Scholz H, Haverich A, Hirt S, Doring
V, Kalmar P, Hoppner W. Increased messenger RNA level of the inhibitory G protein
alpha subunit Gi alpha-2 in human end-stage heart failure. Circ Res 70: 688 – 696, 1992.
151. Evans BA, Sato M, Sarwar M, Hutchinson DS, Summers RJ. Ligand-directed signaling
at beta-adrenoceptors. Br J Pharmacol 159: 1022–1038, 2010.
152. Evans MI, Dougan MB, Moawad AH, Evans WJ, Bryant GG, Greenwood FC. Ripening
of the human cervix with porcine ovarian relaxin. Am J Obstet Gynecol 147: 410 – 414,
1983.
153. Evans RM, Barish GD, Wang YX. PPARs and the complex journey to obesity. Nat Med
10: 355–361, 2004.
154. Failli P, Nistri S, Quattrone S, Mazzetti L, Bigazzi M, Sacchi TB, Bani D. Relaxin
up-regulates inducible nitric oxide synthase expression and nitric oxide generation in
rat coronary endothelial cells. FASEB J 16: 252–254, 2002.
155. Familari M, Vu D, Parry LJ. Regulation of Rxfp2 (Lgr8) expression in the mouse fetal
kidney by the transcription factor Pod1 (Tcf 21). Ann NY Acad Sci 1160: 317–319,
2009.
156. Fei DT, Gross MC, Lofgren JL, Mora-Worms M, Chen AB. Cyclic AMP response to
recombinant human relaxin by cultured human endometrial cells–a specific and high
throughput in vitro bioassay. Biochem Biophys Res Commun 170: 214 –222, 1990.
157. Feng S, Agoulnik IU, Bogatcheva NV, Kamat AA, Kwabi-Addo B, Li R, Ayala G, Ittmann
MM, Agoulnik AI. Relaxin promotes prostate cancer progression. Clin Cancer Res 13:
1695–1702, 2007.
158. Feng S, Bogatcheva NV, Kamat AA, Truong A, Agoulnik AI. Endocrine effects of
relaxin overexpression in mice. Endocrinology 147: 407– 414, 2006.
159. Feng S, Bogatcheva NV, Truong A, Engel W, Adham IM, Agoulnik AI. Overexpression
of insulin-like 3 does not prevent cryptorchidism in GNRHR or HOXA10 deficient
mice. J Urol 176: 399 – 404, 2006.
160. Ferlin A, Menegazzo M, Gianesello L, Selice R, Foresta C. Effect of relaxin on human
sperm functions and fertilizing ability. J Androl 2011.
161. Ferlin A, Pepe A, Gianesello L, Garolla A, Feng S, Facciolli A, Morello R, Agoulnik AI,
Foresta C. New roles for INSL3 in adults. Ann NY Acad Sci 1160: 215–218, 2009.
162. Ferlin A, Pepe A, Gianesello L, Garolla A, Feng S, Giannini S, Zaccolo M, Facciolli A,
Morello R, Agoulnik AI, Foresta C. Mutations in the insulin-like factor 3 receptor are
associated with osteoporosis. J Bone Miner Res 23: 683– 693, 2008.
163. Fernandez-Patron C, Radomski MW, Davidge ST. Vascular matrix metalloprotei-
nase-2 cleaves big endothelin-1 yielding a novel vasoconstrictor. Circ Res 85: 906 –911,
1999.
164. Fernandez-Patron C, Stewart KG, Zhang Y, Koivunen E, Radomski MW, Davidge ST.
Vascular matrix metalloproteinase-2-dependent cleavage of calcitonin gene-related
peptide promotes vasoconstriction. Circ Res 87: 670 – 676, 2000.
165. Fevold HL, Hisaw FL, Meyer RK. The relaxative hormone of the corpus luteum. Its
purification and concentration. J Am Chem Soc 52: 3340 –3348, 1930.
166. Fields PA, Larkin LH. Purification and immunohistochemical localization of relaxin in
the human term placenta. J Clin Endocrinol Metab 52: 79 – 85, 1981.
167. Filonzi M, Cardoso LC, Pimenta MT, Queiroz DB, Avellar MC, Porto CS, Lazari MF.
Relaxin family peptide receptors Rxfp1 and Rxfp2: mapping of the mRNA and protein
distribution in the reproductive tract of the male rat. Reprod Biol Endocrinol 5: 29,
2007.
168. Finn CA, Pope MD, Milligan SR. Relaxin and decidualization in mice: a reappraisal. Biol
Reprod 55: 1415–1418, 1996.
169. Fisher C, Berry C, Blue L, Morton JJ, McMurray J. NH
2
-terminal pro B type natriuretic
peptide, but not the new putative cardiac hormone relaxin, predicts prognosis in
patients with chronic heart failure. Heart 89: 879 – 881, 2003.
170. Fisher C, MacLean M, Morecroft I, Seed A, Johnston F, Hillier C, McMurray J. Is the
pregnancy hormone relaxin also a vasodilator peptide secreted by the heart? Circula-
tion 106: 292–295, 2002.
171. Fisher CJ. Relaxin: a new cardiovascular hormone in humans? Comparative potency
and mechanisms of action. In: Department of Medicine and Therapeutics . Glasgow:
Glasgow, 2009, p. 216.
172. Foresta C, Bettella A, Vinanzi C, Dabrilli P, Meriggiola MC, Garolla A, Ferlin A. A novel
circulating hormone of testis origin in humans. J Clin Endocrinol Metab 89: 5952–5958,
2004.
173. Foresta C, Ferlin A. Role of INSL3 and LGR8 in cryptorchidism and testicular func-
tions. Reprod Biomed Online 9: 294 –298, 2004.
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
469Physiol Rev VOL 93 JANUARY 2013 www.prv.org
174. Frieden EH, Hisaw FL. The biochemistry of relaxin. Rec Prog Horm Res 8: 333–378,
1953.
175. Fu P, Shen PJ, Zhao CX, Scott DJ, Samuel CS, Wade JD, Tregear GW, Bathgate RA,
Gundlach AL. Leucine-rich repeat-containing G-protein-coupled receptor 8 in ma-
ture glomeruli of developing and adult rat kidney and inhibition by insulin-like pep-
tide-3 of glomerular cell proliferation. J Endocrinol 189: 397– 408, 2006.
176. Funato N, Ohyama K, Kuroda T, Nakamura M. Basic helix-loop-helix transcription
factor epicardin/capsulin/Pod-1 suppresses differentiation by negative regulation of
transcription. J Biol Chem 278: 7486 –7493, 2003.
177. Gafvels ME, Caid M, Britt D, Jackson CL, Patterson D, Strauss JFI. Cloning of a cDNA
encoding a putative human very low density lipoprotein/apolipoprotein E receptor
and assignment of the gene to chromosome 9pter-p23. Somat Cell Mol Genet 19:
557–569, 1993.
178. Gambineri A, Patton L, De Iasio R, Palladoro F, Pagotto U, Pasquali R. Insulin-like
factor 3: a new circulating hormone related to luteinizing hormone-dependent ovar-
ian hyperandrogenism in the polycystic ovary syndrome. J Clin Endocrinol Metab 92:
2066 –2073, 2007.
179. Gandley RE, Conrad KP, McLaughlin MK. Endothelin and nitric oxide mediate re-
duced myogenic reactivity of small renal arteries from pregnant rats. Am J Physiol Regul
Integr Comp Physiol 280: R1–R7, 2001.
180. Ganesan A, Klonisch T, McGuane JT, Feng S, Agoulnik AI, Parry LJ. Normal prostate
morphology in relaxin-mutant mice. Reprod Fertil Dev 21: 440 – 450, 2009.
181. Garber SL, Mirochnik Y, Brecklin CS, Unemori EN, Singh AK, Slobodskoy L, Grove
BH, Arruda JA, Dunea G. Relaxin decreases renal interstitial fibrosis and slows pro-
gression of renal disease. Kidney Int 59: 876 – 882, 2001.
182. Ghosh D, Stewart DR, Nayak NR, Lasley BL, Overstreet JW, Hendrickx AG, Sengupta
J. Serum concentrations of oestradiol-17beta, progesterone, relaxin and chorionic
gonadotrophin during blastocyst implantation in natural pregnancy cycle and in em-
bryo transfer cycle in the rhesus monkey. Human Reprod 12: 914 –920, 1997.
183. Gill WB, Schumacher GF, Bibbo M, Straus FH 2nd, Schoenberg HW. Association of
diethylstilbestrol exposure in utero with cryptorchidism, testicular hypoplasia and
semen abnormalities. J Urol 122: 36 –39, 1979.
184. Gnessi L, Basciani S, Mariani S, Arizzi M, Spera G, Wang C, Bondjers C, Karlsson L,
Betsholtz C. Leydig cell loss and spermatogenic arrest in platelet-derived growth
factor (PDGF)-A-deficient mice. J Cell Biol 149: 1019 –1026, 2000.
185. Goldsmith LT, Weiss G, Palejwala S, Plant TM, Wojtczuk A, Lambert WC, Ammur N,
Heller D, Skurnick JH, Edwards D, Cole DM. Relaxin regulation of endometrial struc-
ture and function in the rhesus monkey. Proc Natl Acad Sci USA 101: 4685– 4689, 2004.
186. Gorlov IP, Kamat A, Bogatcheva NV, Jones E, Lamb DJ, Truong A, Bishop CE, Mc-
Elreavey K, Agoulnik AI. Mutations of the GREAT gene cause cryptorchidism. Hum
Mol Genet 11: 2309 –2318, 2002.
187. Goto M, Swanson LW, Canteras NS. Connections of the nucleus incertus. J Comp
Neurol 438: 86 –122, 2001.
188. Graves PN, Tomer Y, Davies TF. Cloning and sequencing of a 1.3 KB variant of human
thyrotropin receptor mRNA lacking the transmembrane domain. Biochem Biophys Res
Commun 187: 1135–1143, 1992.
189. Gundlach AL, Ma S, Sang Q, Shen PJ, Piccenna L, Sedaghat K, Smith CM, Bathgate RA,
Lawrence AJ, Tregear GW, Wade JD, Finkelstein DI, Bonaventure P, Liu C, Loven-
berg TW, Sutton SW. Relaxin family peptides and receptors in mammalian brain. Ann
NY Acad Sci 1160: 226 –235, 2009.
190. Gunnersen JM, Crawford RJ, Tregear GW. Expression of the relaxin gene in rat
tissues. Mol Cell Endo 110: 55– 64, 1995.
191. Gunnersen JM, Fu P, Roche PJ, Tregear GW. Expression of human relaxin genes:
characterization of a novel alternatively-spliced human relaxin mRNA species. Mol
Cell Endocrinol 118: 85–94, 1996.
192. Haley J, Hudson P, Scanlon D, John M, Cronk M, Shine J, Tregear G, Niall H. Porcine
relaxin: molecular cloning and cDNA structure. DNA 1: 155–162, 1982.
193. Hall K. Further notes on the action of oestrone and relaxin on the pelvis of the spayed
mouse, including a single-dose test of potency of relaxin. J Endocrinol 5: 314 –321,
1948.
194. Hall K. Modification by relaxin of the response of the reproductive tract of mice to
oestradiol and progesterone. J Endocrinol 20: 355–364, 1960.
195. Hall K, Newton WH. The effect of oestrone and relaxin on the X-ray appearance of
the pelvis of the mouse. J Physiol 106: 18 –27, 1947.
196. Halls ML. Constitutive formation of an RXFP1-signalosome: a novel paradigm in
GPCR function and regulation. Br J Pharmacol 165: 1644 –1658, 2012.
197. Halls ML, Bathgate RA, Summers RJ. Comparison of signaling pathways activated by
the relaxin family peptide receptors, RXFP1 and RXFP2, using reporter genes. J
Pharmacol Exp Ther 320: 281–290, 2007.
198. Halls ML, Bathgate RA, Summers RJ. Relaxin family peptide receptors RXFP1 and
RXFP2 modulate cAMP signaling by distinct mechanisms. Mol Pharmacol 70: 214 –226,
2006.
199. Halls ML, Bond CP, Sudo S, Kumagai J, Ferraro T, Layfield S, Bathgate RA, Summers
RJ. Multiple binding sites revealed by interaction of relaxin family peptides with native
and chimeric relaxin family peptide receptors 1 and 2 (LGR7 and LGR8). J Pharmacol
Exp Ther 313: 677– 687, 2005.
200. Halls ML, Cooper DM. Regulation by Ca
2
-signaling pathways of adenylyl cyclases.
Cold Spring Harb Perspect Biol 3: a004143, 2011.
201. Halls ML, Cooper DM. Sub-picomolar relaxin signaling by a pre-assembled RXFP1,
AKAP79, AC2, beta-arrestin 2, PDE4D3 complex. EMBO J 29: 2772–2787, 2010.
202. Halls ML, Hewitson TD, Moore XL, Du XJ, Bathgate RA, Summers RJ. Relaxin acti-
vates multiple cAMP signaling pathway profiles in different target cells. Ann NY Acad Sci
1160: 108 –111, 2009.
203. Halls ML, van der Westhuizen ET, Bathgate RA, Summers RJ. Relaxin family peptide
receptors–former orphans reunite with their parent ligands to activate multiple sig-
naling pathways. Br J Pharmacol 150: 677– 691, 2007.
204. Halls ML, van der Westhuizen ET, Wade JD, Evans BA, Bathgate RA, Summers RJ.
Relaxin family peptide receptor (RXFP1) coupling to G(alpha)i3 involves the COOH-
terminal Arg752 and localization within membrane Raft Microdomains. Mol Pharmacol
75: 415– 428, 2009.
205. Hamolsky M, Sparrow RC. Influence of relaxin on mammary development in sexually
immature female rats. Proc Soc Exp Biol Med 60: 8, 1945.
206. Han X, Habuchi Y, Giles WR. Relaxin increases heart rate by modulating calcium
current in cardiac pacemaker cells. Circ Res 74: 537–541, 1994.
207. Handoko ML, de Man FS, Vonk-Noordegraaf A. The rise and fall of endothelin recep-
tor antagonists in congestive heart failure. Eur Respir J 37: 484 – 485, 2011.
208. Hansell DJ, Bryant GG, Greenwood FC. Expression of the human relaxin H1 gene in
the decidua, trophoblast, and prostate. J Clin Endocrinol Metab 72: 899 –904, 1991.
209. Harris R, Finlayson C, Weiss J, Fisher L, Hurley L, Barrett T, Emge D, Bathgate R,
Agoulnik A, Jameson J. A missense mutation in LRR8 of RXFP2 is associated with
cryptorchidism. Mammalian Genome 21: 442– 449, 2010.
210. Hartley BJ, Scott DJ, Callander GE, Wilkinson TN, Ganella DE, Kong CK, Layfield S,
Ferraro T, Petrie EJ, Bathgate RA. Resolving the unconventional mechanisms under-
lying RXFP1 and RXFP2 receptor function. Ann NY Acad Sci 1160: 67–73, 2009.
211. Haugaard-Jonsson LM, Hossain MA, Daly NL, Bathgate RA, Wade JD, Craik DJ,
Rosengren KJ. Structure of the R3/I5 chimeric relaxin peptide, a selective GPCR135
and GPCR142 agonist. J Biol Chem 283: 23811–23818, 2008.
212. Haugaard-Jonsson LM, Hossain MA, Daly NL, Craik DJ, Wade JD, Rosengren KJ.
Structure of human insulin-like peptide 5 and characterization of conserved hydrogen
bonds and electrostatic interactions within the relaxin framework. Biochem J 419:
619 – 627, 2009.
213. Haugaard-Kedstrom LM, Shabanpoor F, Hossain MA, Clark RJ, Ryan PJ, Craik DJ,
Gundlach AL, Wade JD, Bathgate RA, Rosengren KJ. Design, synthesis, and charac-
terization of a single-chain peptide antagonist for the relaxin-3 receptor RXFP3. JAm
Chem Soc 133: 4965– 4974, 2011.
214. Hayes ES, Curnow EC, Trounson AO, Danielson LA, Unemori EN. Implantation and
pregnancy following in vitro fertilization and the effect of recombinant human relaxin
administration in Macaca fascicularis.Biol Reprod 71: 1591–1597, 2004.
BATHGATE ET AL.
470 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
215. Heeg MHJ, Koziolek MJ, Vasko R, Schaefer L, Sharma K, Muller GA, Strutz F. The
antifibrotic effects of relaxin in human renal fibroblasts are mediated in part by inhi-
bition of the Smad2 pathway. Kidney Int 68: 96 –109, 2005.
216. Heng K, Ivell R, Wagaarachchi P, Anand-Ivell R. Relaxin signaling in primary cultures of
human myometrial cells. Mol Hum Reprod 14: 603– 611, 2008.
217. Herrick-Davis K, Egan CT, Teitler M. Activating mutations of the serotonin 5-HT
2C
receptor. J Neurochem 69: 1138 –1144, 1997.
218. Herz J, Hamann U, Rogne S, Myklebost O, Gausepohl H, Stanley KK. Surface location
and high affinity for calcium of a 500kD live membrane protein closely related to the
LDL-receptor suggest a physiological role as a lipoprotein receptor. EMBO J 7: 4119 –
4127, 1988.
219. Hida T, Takahashi E, Shikata K, Hirohashi T, Sawai T, Seiki T, Tanaka H, Kawai T, Ito
O, Arai T, Yokoi A, Hirakawa T, Ogura H, Nagasu T, Miyamoto N, Kuromitsu J.
Chronic intracerebroventricular administration of relaxin-3 increases body weight in
rats. J Recept Signal Transduct Res 26: 147–158, 2006.
220. Hisaw FL. Experimental relaxation of the pubic ligament of the guinea pig. Proc Soc Exp
Biol Med 23: 661– 663, 1926.
221. Hisaw FL. Experimental relaxation of the symphysis pubis of the guinea pig. Anat Rec
37: 126, 1927.
222. Hisaw FL, Hisaw FLJ, Dawson AB. Effects of relaxin on the endothelium of endome-
trial blood vessels in monkeys (Macaca mulatta). Endocrinology 81: 375–385, 1967.
223. Hjalm G, Murray E, Crumley G, Harazim W, Lundgren S, Onyango I, Ek B, Larsson M,
Juhlin C, Hellman P, Davis H, Akerstrom G, Rask L, Morse B. Cloning and sequencing
of human GP330, a Ca
2
-binding recptor with potential intracellular signaling prop-
erties. Eur J Biochem 239: 132–137, 1996.
224. Ho TY, Yan W, Bagnell CA. Relaxin-induced matrix metalloproteinase-9 expression is
associated with activation of the NF-
B pathway in human THP-1 cells. J Leukoc Biol
81: 1303–1310, 2007.
225. Hombach-Klonisch S, Bialek J, Radestock Y, Truong A, Agoulnik AI, Fiebig B, Willing
C, Weber E, Hoang-Vu C, Klonisch T. INSL3 has tumor-promoting activity in thyroid
cancer. Int J Cancer 127: 521–531, 2010.
226. Hombach-Klonisch S, Bialek J, Trojanowicz B, Weber E, Holzhausen HJ, Silvertown
JD, Summerlee AJ, Dralle H, Hoang-Vu C, Klonisch T. Relaxin enhances the oncogenic
potential of human thyroid carcinoma cells. Am J Pathol 169: 617– 632, 2006.
227. Hombach-Klonisch S, Buchmann J, Sarun S, Fischer B, Klonisch T. Relaxin-like factor
(RLF) is differentially expressed in the normal and neoplastic human mammary gland.
Cancer 89: 2161–2168, 2000.
228. Hombach-Klonisch S, Hoang-Vu C, Kehlen A, Hinze R, Holzhausen HJ, Weber E,
Fischer B, Dralle H, Klonisch T. INSL-3 is expressed in human hyperplastic and
neoplastic thyrocytes. Int J Oncol 22: 993–1001, 2003.
229. Hombach-Klonisch S, Schon J, Kehlen A, Blottner S, Klonisch T. Seasonal expression
of INSL3 and LGR8/INSL3 receptor transcripts indicates variable differentiation of
Leydig cells in the roe deer testis. Biol Reprod 71: 1079 –1087, 2004.
230. Hombach-Klonisch S, Seeger S, Tscheudschilsuren G, Buchmann J, Huppertz B,
Seliger G, Fischer B, Klonisch T. Cellular localization of human relaxin-like factor in the
cyclic endometrium and placenta. Mol Hum Reprod 7: 349 –356, 2001.
231. Hopkins EJ, Layfield S, Ferraro T, Bathgate RA, Gooley PR. The NMR solution struc-
ture of the relaxin (RXFP1) receptor lipoprotein receptor class A module and iden-
tification of key residues in the NH
2
-terminal region of the module that mediate
receptor activation. J Biol Chem 282: 4172– 4184, 2007.
232. Hossain MA, Rosengren KJ, Haugaard-Jonsson LM, Zhang S, Layfield S, Ferraro T,
Daly NL, Tregear GW, Wade JD, Bathgate RA. The A-chain of human relaxin family
peptides has distinct roles in the binding and activation of the different relaxin family
peptide receptors. J Biol Chem 283: 17287–17297, 2008.
233. Hossain MA, Rosengren KJ, Samuel CS, Shabanpoor F, Chan LJ, Bathgate RA, Wade
JD. The minimal active structure of human relaxin-2. J Biol Chem 286: 37555–37565,
2011.
234. Houslay MD. Underpinning compartmentalised cAMP signaling through targeted
cAMP breakdown. Trends Biochem Sci 35: 91–100, 2010.
235. Hsu CJ, McCormack SM, Sanborn BM. The effect of relaxin on cyclic adenosine
3=,5=-monophosphate concentrations in rat myometrial cells in culture. Endocrinology
116: 2029 –2035, 1985.
236. Hsu SY. Cloning of two novel mammalian paralogs of relaxin/insulin family proteins
and their expression in testis and kidney. Mol Endocrinol 13: 2163–2174, 1999.
237. Hsu SY, Kudo M, Chen T, Nakabayashi K, Bhalla A, van der Spek PJ, van Duin M,
Hsueh AJ. The three subfamilies of leucine-rich repeat-containing G protein-coupled
receptors (LGR): identification of LGR6 and LGR7 and the signaling mechanism for
LGR7. Mol Endocrinol 14: 1257–1271, 2000.
238. Hsu SY, Liang SG, Hsueh AJ. Characterization of two LGR genes homologous to
gonadotropin and thyrotropin receptors with extracellular leucine-rich repeats and a
G protein-coupled, seven-transmembrane region. Mol Endocrinol 12: 1830 –1845,
1998.
239. Hsu SY, Nakabayashi K, Nishi S, Kumagai J, Kudo M, Sherwood OD, Hsueh AJ.
Activation of orphan receptors by the hormone relaxin. Science 295: 671– 674, 2002.
240. Hsu SY, Semyonov J, Park JI, Chang CL. Evolution of the signaling system in relaxin-
family peptides. Ann NY Acad Sci 1041: 520 –529, 2005.
241. Hsu SYT. New insights into the evolution of the relaxin-LGR signaling system. Trends
Endocrinol Metab 14: 303–309, 2003.
242. Huang CJ, Li Y, Anderson LL. Relaxin and estrogen synergistically accelerate growth
and development in the uterine cervix of prepubertal pigs. Anim Reprod Sci 46: 149 –
158, 1997.
243. Huang JR, Tseng L, Bischof P, Janne OA. Regulation of prolactin production by pro-
gestin, estrogen, and relaxin in human endometrial stromal cells. Endocrinology 121:
2011–2017, 1987.
244. Hudson P, Haley J, Cronk M, Shine J, Niall H. Molecular cloning and characterization
of cDNA sequences coding for rat relaxin. Nature 291: 127–131, 1981.
245. Hudson P, Haley J, John M, Cronk M, Crawford R, Haralambidis J, Tregear G, Shine J,
Niall H. Structure of a genomic clone encoding biologically active human relaxin.
Nature 301: 628 – 631, 1983.
246. Hudson P, John M, Crawford R, Haralambidis J, Scanlon D, Gorman J, Tregear G,
Shine J, Niall H. Relaxin gene expression in human ovaries and the predicted structure
of a human preprorelaxin by analysis of cDNA clones. EMBO J 3: 2333–2339, 1984.
247. Hughes SCC, Mason HD, Franks S, Holly JMP. The insulin-like growth factors (IGFs)
in follicular fluid are predominantly bound in the ternary complex. J Endocrinol 155:
R1–R4, 1997.
248. Hurley WL, Doane RM, O’Day-Bowman MB, Winn RJ, Mojonnier LE, Sherwood OD.
Effect of relaxin on mammary development in ovariectomized pregnant gilts. Endo-
crinology 128: 1285–1290, 1991.
249. Huson DH, Richter DC, Rausch C, Dezulian T, Franz M, Rupp R. Dendroscope: an
interactive viewer for large phylogenetic trees. BMC Bioinformatics 8: 460, 2007.
250. Hutson JM, Hasthorpe S, Heyns CF. Anatomical and functional aspects of testicular
descent and cryptorchidism. Endocr Rev 18: 259 –280, 1997.
251. Hwang JJ, Lee AB, Fields PA, Haab LM, Mojonnier LE, Sherwood OD. Monoclonal
antibodies specific for rat relaxin. V. Passive immunization with monoclonal antibodies
throughout the second half of pregnancy disrupts development of the mammary
apparatus and, hence, lactational performance in rats. Endocrinology 129: 3034 –3042,
1991.
252. Hwang JJ, Sherwood OD. Monoclonal antibodies specific for rat relaxin. III. Passive
immunization with monoclonal antibodies throughout the second half of pregnancy
reduces cervical growth and extensibility in intact rats. Endocrinology 123: 2486 –2490,
1988.
253. Irving-Rodgers HF, Bathgate RA, Ivell R, Domagalski R, Rodgers RJ. Dynamic changes
in the expression of relaxin-like factor (INSL3), cholesterol side-chain cleavage cyto-
chrome p450, and 3beta-hydroxysteroid dehydrogenase in bovine ovarian follicles
during growth and atresia. Biol Reprod 66: 934 –943, 2002.
254. Ivell R, Balvers M, Domagalski R, Ungefroren H, Hunt N, Schulze W. Relaxin-like
factor: a highly specific and constitutive new marker for Leydig cells in the human
testis. Mol Hum Reprod 3: 459 – 466, 1997.
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
471Physiol Rev VOL 93 JANUARY 2013 www.prv.org
255. Ivell R, Balvers M, Pohnke Y, Telgmann R, Bartsch O, Milde-Langosch K, Bamberger
AM, Einspanier A. Immunoexpression of the relaxin receptor LGR7 in breast and
uterine tissues of humans and primates. Reprod Biol Endocrinol 1: 114, 2003.
256. Ivell R, Bathgate RA. Reproductive biology of the relaxin-like factor (RLF/INSL3). Biol
Reprod 67: 699 –705, 2002.
257. Ivell R, Hartung S. The molecular basis of cryptorchidism. Mol Hum Reprod 9: 175–
181, 2003.
258. Ivell R, Hunt N, Khan-Dawood F, Dawood MY. Expression of the human relaxin gene
in the corpus luteum of the menstrual cycle and in the prostate. Mol Cell Endocrinol 66:
251–255, 1989.
259. Ivell R, Kotula-Balak M, Glynn D, Heng K, Anand-Ivell R. Relaxin family peptides in the
male reproductive system–a critical appraisal. Mol Hum Reprod 17: 71– 84, 2011.
260. Jaspers S, Lok S, Lofton-Day C, Haldeman B, Whitmore T, Foley K, Conklin D.
(Editors). The Genomics of Insulin 5 . Dordrecht, Germany: Kluwer, 2000.
261. Jennes L, Beckman WC, Stumpf WE, Grzanna R. Anatomical relationships of sero-
toninergic and noradrenalinergic projections with the GnRH system in septum and
hypothalamus. Exp Brain Res 46: 331–338, 1982.
262. Jeyabalan A, Kerchner LJ, Fisher MC, McGuane JT, Doty KD, Conrad KP. Matrix
metalloproteinase-2 activity, protein, mRNA, and tissue inhibitors in small arteries
from pregnant and relaxin-treated nonpregnant rats. J Appl Physiol 100: 1955–1963,
2006.
263. Jeyabalan A, Novak J, Danielson LA, Kerchner LJ, Opett SL, Conrad KP. Essential role
for vascular gelatinase activity in relaxin-induced renal vasodilation, hyperfiltration,
and reduced myogenic reactivity of small arteries. Circ Res 93: 1249 –1257, 2003.
264. Jeyabalan A, Novak J, Doty KD, Matthews J, Fisher MC, Kerchner LJ, Conrad KP.
Vascular matrix metalloproteinase-9 mediates the inhibition of myogenic reactivity in
small arteries isolated from rats after short-term administration of relaxin. Endocrinol-
ogy 148: 189 –197, 2007.
265. Jeyabalan A, Shroff SG, Novak J, Conrad KP. The vascular actions of relaxin. Adv Exp
Med Biol 612: 65– 87, 2007.
266. Johnson KJ, Robbins AK, Wang Y, McCahan SM, Chacko JK, Barthold JS. Insulin-like 3
exposure of the fetal rat gubernaculum modulates expression of genes involved in
neural pathways. Biol Reprod 83: 774 –782, 2010.
267. Johnson MR, Abdalla H, Allman AC, Wren ME, Kirkland A, Lightman SL. Relaxin levels
in ovum donation pregnancies. Fertil Steril 56: 59 – 61, 1991.
268. Johnson MR, Brooks AA, Steer PJ. The role of relaxin in the pregnancy associated
reduction in plasma osmolality. Hum Reprod 11: 1105–1108, 1996.
269. Johnson MR, Okokon E, Collins WP, Sharma V, Lightman SL. The effect of human
chorionic gonadotropin and pregnancy on the circulating level of relaxin. J Clin Endo-
crinol Metab 72: 1042–1047, 1991.
270. Kaftanovskaya EM, Feng S, Huang Z, Tan Y, Barbara AM, Kaur S, Truong A, Gorlov
IP, Agoulnik AI. Suppression of insulin-like3 receptor reveals the role of beta-
catenin and Notch signaling in gubernaculum development. Mol Endocrinol 25:
170 –183, 2011.
271. Kakouris H, Eddie LW, Summers RJ. Cardiac effects of relaxin in rats. Lancet 339:
1076 –1078, 1992.
272. Kalbag SS, Roginsky MS, Jelveh Z, Sulimovici S. Phorbol ester, prolactin, and relaxin
cause translocation of protein kinase C from cytosol to membranes in human endo-
metrial cells. Biochim Biophys Acta 1094: 85–91, 1991.
273. Kamat AA, Feng S, Agoulnik IU, Kheradmand F, Bogatcheva NV, Coffey D, Sood AK,
Agoulnik AI. The role of relaxin in endometrial cancer. Cancer Biol Ther 5: 71–77,
2006.
274. Kamat AA, Feng S, Bogatcheva NV, Truong A, Bishop CE, Agoulnik AI. Genetic
targeting of relaxin and insulin-like factor 3 receptors in mice. Endocrinology 145:
4712– 4720, 2004.
275. Kasik J, Muglia L, Stephan DA, Menon RK. Identification, chromosomal mapping, and
partial characterization of mouse Insl6: a new member of the insulin family. Endocri-
nology 141: 458 – 461, 2000.
276. Kawamura K, Kumagai J, Sudo S, Chun SY, Pisarska M, Morita H, Toppari J, Fu P,
Wade JD, Bathgate RA, Hsueh AJ. Paracrine regulation of mammalian oocyte matu-
ration and male germ cell survival. Proc Natl Acad Sci USA 101: 7323–7328, 2004.
277. Kenakin T, Miller LJ. Seven transmembrane receptors as shapeshifting proteins: the
impact of allosteric modulation and functional selectivity on new drug discovery.
Pharmacol Rev 62: 265–304, 2010.
278. Kenakin TP. The secret lives of GPCRs. Drug Discov Today 8: 674, 2003.
279. Kerchner LJ, Novak J, Hanley-Yanez K, Doty KD, Danielson LA, Conrad KP. Evidence
against the hypothesis that endothelial endothelin B receptor expression is regulated
by relaxin and pregnancy. Endocrinology 146: 2791–2797, 2005.
280. Kern A, Agoulnik AI, Bryant-Greenwood GD. The low-density lipoprotein class A
module of the relaxin receptor (leucine-rich repeat containing G-protein coupled
receptor 7): its role in signaling and trafficking to the cell membrane. Endocrinology
148: 1181–1194, 2007.
281. Kern A, Bryant-Greenwood GD. Characterization of relaxin receptor (RXFP1) de-
sensitization and internalization in primary human decidual cells and RXFP1-trans-
fected HEK293 cells. Endocrinology 150: 2419 –2428, 2009.
282. Kern A, Hubbard D, Amano A, Bryant-Greenwood GD. Cloning, expression, and
functional characterization of relaxin receptor (leucine-rich repeat-containing g pro-
tein-coupled receptor 7) splice variants from human fetal membranes. Endocrinology
149: 1277–1294, 2008.
283. Khan-Dawood FS, Goldsmith LT, Weiss G, Dawood MY. Human corpus luteum
secretion of relaxin, oxytocin, and progesterone. J Clin Endocrinol Metab 68: 627– 631,
1989.
284. Kjelsberg MA, Cotecchia S, Ostrowski J, Caron MG, Lefkowitz RJ. Constitutive acti-
vation of the alpha 1B-adrenergic receptor by all amino acid substitutions at a single
site. Evidence for a region which constrains receptor activation. J Biol Chem 267:
1430 –1433, 1992.
285. Klonisch T, Hombach-Klonisch S, Buchmann J, Fischer B, Bergmann M, Steger K.
Relaxin-like factor expression in a human ovarian Sertoli-Leydig cell tumor. Fertil Steril
72: 546 –548, 1999.
286. Klonisch T, Muller-Huesmann H, Riedel M, Kehlen A, Bialek J, Radestock Y, Hol-
zhausen HJ, Steger K, Ludwig M, Weidner W, Hoang-Vu C, Hombach-Klonisch S.
INSL3 in the benign hyperplastic and neoplastic human prostate gland. Int J Oncol 27:
307–315, 2005.
287. Koay ES, Bagnell CA, Bryant GG, Lord SB, Cruz AC, Larkin LH. Immunocytochemical
localization of relaxin in human decidua and placenta. J Clin Endocrinol Metab 60:
859 – 863, 1985.
288. Kohsaka T, Min G, Lukas G, Trupin S, Campbell ET, Sherwood OD. Identification of
specific relaxin-binding cells in the human female. Biol Reprod 59: 991–999, 1998.
289. Koman A, Cazaubon S, Couraud PO, Ullrich A, Strosberg AD. Molecular character-
ization and in vitro biological activity of placentin, a new member of the insulin gene
family. J Biol Chem 271: 20238 –20241, 1996.
290. Kompa AR, Samuel CS, Summers RJ. Inotropic responses to human gene 2 (B29)
relaxin in a rat model of myocardial infarction (MI): effect of pertussis toxin. Br J
Pharmacol 137: 710 –718, 2002.
291. Kong RC, Shilling PJ, Lobb DK, Gooley PR, Bathgate RA. Membrane receptors: struc-
ture and function of the relaxin family peptide receptors. Mol Cell Endocrinol 320:
1–15, 2010.
292. Koskimies P, Suvanto M, Nokkala E, Huhtaniemi IT, McLuskey A, Themmen AP,
Poutanen M. Female mice carrying a ubiquitin promoter-Insl3 transgene have de-
scended ovaries and inguinal hernias but normal fertility. Mol Cell Endocrinol 206:
159 –166, 2003.
293. Krajnc-Franken MA, van Disseldorp AJ, Koenders JE, Mosselman S, van Duin M,
Gossen JA. Impaired nipple development and parturition in LGR7 knockout mice. Mol
Cell Biol 24: 687– 696, 2004.
294. Kramer SM, Gibson UE, Fendly BM, Mohler MA, Drolet DW, Johnston PD. Increase
in cyclic AMP levels by relaxin in newborn rhesus monkey uterus cell culture. In Vitro
Cell Dev Biol 26: 647– 656, 1990.
295. Krantz JC, Bryant HH, Carr CJ. The action of aqueous corpus luteum extract upon
uterine activity. Surg Gynecol Obstet 90: 372–375, 1950.
BATHGATE ET AL.
472 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
296. Krum H, Teerlink JR. Medical therapy for chronic heart failure. Lancet 378: 713–721,
2011.
297. Krusche CA, Kroll T, Beier HM, Classen-Linke I. Expression of leucine-rich repeat-
containing G-protein-coupled receptors in the human cyclic endometrium. Fertil Steril
87: 1428 –1437, 2007.
298. Kubota Y, Temelcos C, Bathgate RA, Smith KJ, Scott D, Zhao C, Hutson JM. The role
of insulin 3, testosterone, Mullerian inhibiting substance and relaxin in rat gubernacu-
lar growth. Mol Hum Reprod 8: 900 –905, 2002.
299. Kuei C, Sutton S, Bonaventure P, Pudiak C, Shelton J, Zhu J, Nepomuceno D, Wu J,
Chen J, Kamme F, Seierstad M, Hack MD, Bathgate RA, Hossain MA, Wade JD, Atack
J, Lovenberg TW, Liu C. R3(BDelta23 27)R/I5 chimeric peptide, a selective antagonist
for GPCR135 and GPCR142 over relaxin receptor LGR7: in vitro and in vivo charac-
terization. J Biol Chem 282: 25425–25435, 2007.
300. Kuenzi MJ, Sherwood OD. Immunohistochemical localization of specific relaxin-bind-
ing cells in the cervix, mammary glands, and nipples of pregnant rats. Endocrinology
136: 1367–1373, 1995.
301. Kumagai J, Hsu SY, Matsumi H, Roh JS, Fu P, Wade JD, Bathgate RA, Hsueh AJ.
INSL3/Leydig insulin-like peptide activates the LGR8 receptor important in testis
descent. J Biol Chem 277: 31283–31286, 2002.
302. Kuznetsova L, Plesneva S, Derjabina N, Omeljaniuk E, Pertseva M. On the mechanism
of relaxin action: the involvement of adenylyl cyclase signaling system. Regul Pept 80:
33–39, 1999.
303. Laurent A, Rouillac C, Delezoide AL, Giovangrandi Y, Vekemans M, Bellet D, Abitbol
M, Vidaud M. Insulin-like 4 (INSL4) gene expression in human embryonic and tropho-
blastic tissues. Mol Reprod Dev 51: 123–129, 1998.
304. Lee AB, Hwang JJ, Haab LM, Fields PA, Sherwood OD. Monoclonal antibodies specific
for rat relaxin. VI. Passive immunization with monoclonal antibodies throughout the
second half of pregnancy disrupts histological changes associated with cervical soft-
ening at parturition in rats. Endocrinology 130: 2386 –2391, 1992.
305. Lee HY, Zhao S, Fields PA, Sherwood OD. The extent to which relaxin promotes
proliferation and inhibits apoptosis of cervical epithelial and stromal cells is greatest
during late pregnancy in rats. Endocrinology 146: 511–518, 2005.
306. Lein ES, Hawrylycz MJ, Ao N, Ayres M, Bensinger A, Bernard A, Boe AF, Boguski MS,
Brockway KS, Byrnes EJ, Chen L, Chen TM, Chin MC, Chong J, Crook BE, Czaplinska
A, Dang CN, Datta S, Dee NR, Desaki AL, Desta T, Diep E, Dolbeare TA, Donelan MJ,
Dong HW, Dougherty JG, Duncan BJ, Ebbert AJ, Eichele G, Estin LK, Faber C, Facer
BA, Fields R, Fischer SR, Fliss TP, Frensley C, Gates SN, Glattfelder KJ, Halverson KR,
Hart MR, Hohmann JG, Howell MP, Jeung DP, Johnson RA, Karr PT, Kawal R, Kidney
JM, Knapik RH, Kuan CL, Lake JH, Laramee AR, Larsen KD, Lau C, Lemon TA, Liang
AJ, Liu Y, Luong LT, Michaels J, Morgan JJ, Morgan RJ, Mortrud MT, Mosqueda NF, Ng
LL, Ng R, Orta GJ, Overly CC, Pak TH, Parry SE, Pathak SD, Pearson OC, Puchalski
RB, Riley ZL, Rockett HR, Rowland SA, Royall JJ, Ruiz MJ, Sarno NR, Schaffnit K,
Shapovalova NV, Sivisay T, Slaughterbeck CR, Smith SC, Smith KA, Smith BI, Sodt AJ,
Stewart NN, Stumpf KR, Sunkin SM, Sutram M, Tam A, Teemer CD, Thaller C,
Thompson CL, Varnam LR, Visel A, Whitlock RM, Wohnoutka PE, Wolkey CK, Wong
VY, Wood M, Yaylaoglu MB, Young RC, Youngstrom BL, Yuan XF, Zhang B, Zwing-
man TA, Jones AR. Genome-wide atlas of gene expression in the adult mouse brain.
Nature 445: 168 –176, 2007.
307. Lekgabe ED, Kiriazis H, Zhao C, Xu Q, Moore XL, Su Y, Bathgate RA, Du XJ, Samuel
CS. Relaxin reverses cardiac and renal fibrosis in spontaneously hypertensive rats.
Hypertension 46: 412– 418, 2005.
308. Levine RJ, Maynard SE, Qian C, Lim KH, England LJ, Yu KF, Schisterman EF, Thadhani
R, Sachs BP, Epstein FH, Sibai BM, Sukhatme VP, Karumanchi SA. Circulating angio-
genic factors and the risk of preeclampsia. N Engl J Med 350: 672– 683, 2004.
309. Lewis M, Deshpande U, Gizman L, Grove BH, Huang X, Erikson ME, Pickford LB,
Unemori EN. Systemic relaxin administration stimulates angiogenic cytokine expres-
sion and vessel formation in a rat myocardial infarct model. In: Relaxin 2000, Proceed-
ings of the Third International Conference on Relaxin and Related Peptides , edited by
Tregear GW, Ivell R, Bathgate RA, Wade JD. Broome, Australia: Kluwer Academic,
2001, p. 159 –168.
310. Lin F, Otvos L Jr, Kumagai J, Tregear GW, Bathgate RA, Wade JD. Synthetic human
insulin 4 does not activate the G-protein-coupled receptors LGR7 or LGR8. J Pept Sci
10: 257–264, 2004.
311. Lindheimer MD, Barron WM, Davison JM. Osmoregulation of thirst and vasopressin
release in pregnancy. Am J Physiol Renal Fluid Electrolyte Physiol 257: F159 –F169, 1989.
312. Liu C, Chen J, Kuei C, Sutton S, Nepomuceno D, Bonaventure P, Lovenberg TW.
Relaxin-3/insulin-like peptide 5 chimeric peptide, a selective ligand for G protein-
coupled receptor (GPCR)135 and GPCR142 over leucine-rich repeat-containing G
protein-coupled receptor 7. Mol Pharmacol 67: 231–240, 2005.
313. Liu C, Chen J, Sutton S, Roland B, Kuei C, Farmer N, Sillard R, Lovenberg TW.
Identification of relaxin-3/INSL7 as a ligand for GPCR142. J Biol Chem 278: 50765–
50770, 2003.
314. Liu C, Eriste E, Sutton S, Chen J, Roland B, Kuei C, Farmer N, Jornvall H, Sillard R,
Lovenberg TW. Identification of relaxin-3/INSL7 as an endogenous ligand for the
orphan G-protein-coupled receptor GPCR135. J Biol Chem 278: 50754 –50764, 2003.
315. Liu C, Kuei C, Sutton S, Chen J, Bonaventure P, Wu J, Nepomuceno D, Kamme F,
Tran DT, Zhu J, Wilkinson T, Bathgate R, Eriste E, Sillard R, Lovenberg TW. INSL5 is
a high affinity specific agonist for GPCR142 (GPR100). J Biol Chem 280: 292–300,
2005.
316. Loges S, Tinnefeld H, Metzner A, Jucker M, Butzal M, Bruweleit M, Fischer U, Draab
E, Schuch-Farrel AM, Hossfeld DK, Bokemeyer C, Fiedler W. Downregulation of
VEGF-A, STAT5 and AKT in acute myeloid leukemia blasts of patients treated with
SU5416. Leukemia Lymphoma 47: 2601–2609, 2006.
317. Lok S, Johnston DS, Conklin D, Lofton-Day CE, Adams RL, Jelmberg AC, Whitmore
TE, Schrader S, Griswold MD, Jaspers SR. Identification of INSL6, a new member of
the insulin family that is expressed in the testis of the human and rat. Biol Reprod 62:
1593–1599, 2000.
318. Longo M, Jain V, Vedernikov YP, Garfield RE, Saade GR. Effects of recombinant
human relaxin on pregnant rat uterine artery and myometrium in vitro. Am J Obstet
Gynecol 188: 1468 –1474, 2003.
319. Loosfelt H, Misrahi M, Atger M, Salesse R, Vu Hai-Luu Thi M.T, Jolivet A, Guiochon-
Mantel A, Sar S, Jallal B, Garnier J. Cloning and sequencing of porcine LH-hCG recep-
tor cDNA: variants lacking transmembrane domain. Science 245: 525–528, 1989.
320. Loumaye E, De Cooman S, Thomas K. Immunoreactive relaxin-like substance in
human seminal plasma. J Clin Endocrinol Metab 50: 1142–1143, 1980.
321. Loumaye E, Teuwissen B, Thomas K. Characterization of relaxin radioimmuno-
assay using Bolton-Hunter reagent. First results in plasma during pregnancy, in
placenta, corpora lutea and ovarian cysts in woman. Gynecol Obstet Invest 9:
262–267, 1978.
322. Lowndes K, Amano A, Yamamoto SY, Bryant-Greenwood GD. The human relaxin
receptor (LGR7): expression in the fetal membranes and placenta. Placenta 27: 610 –
618, 2006.
323. Lu C, Walker WH, Sun J, Weisz OA, Gibbs RB, Witchel SF, Sperling MA, Menon RK.
Insulin-like peptide 6: characterization of secretory status and posttranslational mod-
ifications. Endocrinology 147: 5611–5623, 2006.
324. Lucas C, Bald LN, Martin MC, Jaffe RB, Drolet DW, Mora WM, Bennett G, Chen AB,
Johnston PD. An enzyme-linked immunosorbent assay to study human relaxin in
human pregnancy and in pregnant rhesus monkeys. J Endocrinol 120: 449 – 457, 1989.
325. Luna JJ, Riesewijk A, Horcajadas JA, de van Os R, Dominguez F, Mosselman S, Pellicer
A, Simon C. Gene expression pattern and immunoreactive protein localization of
LGR7 receptor in human endometrium throughout the menstrual cycle. Mol Hum
Reprod 10: 85–90, 2004.
326. Lundgren S, Hjalm G, Hellman P, Ek B, Juhlin C, Rastad J, Klareskog L, Akerstrom G,
Rask L. A protein involved in calcium sensing of the human parathyroid and placental
cytotrophoblast cells belongs to the LDL-receptor protein superfamily. Exp Cell Res
212: 344 –350, 1994.
327. Luque EH, Munoz de Toro MM, Ramos JG, Rodriguez HA, Sherwood OD. Role of
relaxin and estrogen in the control of eosinophilic invasion and collagen remodeling in
rat cervical tissue at term. Biol Reprod 59: 795– 800, 1998.
328. Luttrell LM, Gesty-Palmer D. Beyond desensitization: physiological relevance of ar-
restin-dependent signaling. Pharmacol Rev 62: 305–330, 2010.
329. Ma S, Bonaventure P, Ferraro T, Shen PJ, Burazin TC, Bathgate RA, Liu C, Tregear
GW, Sutton SW, Gundlach AL. Relaxin-3 in GABA projection neurons of nucleus
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
473Physiol Rev VOL 93 JANUARY 2013 www.prv.org
incertus suggests widespread influence on forebrain circuits via G-protein-coupled
receptor-135 in the rat. Neuroscience 144: 165–190, 2007.
330. Ma S, Olucha-Bordonau FE, Hossain MA, Lin F, Kuei C, Liu C, Wade JD, Sutton SW,
Nunez A, Gundlach AL. Modulation of hippocampal theta oscillations and spatial
memory by relaxin-3 neurons of the nucleus incertus. Learn Mem 16: 730 –742, 2009.
331. Ma S, Roozendaal B, Burazin TC, Tregear GW, McGaugh JL, Gundlach AL. Relaxin
receptor activation in the basolateral amygdala impairs memory consolidation. Eur J
Neurosci 22: 2117–2122, 2005.
332. Ma S, Sang Q, Lanciego JL, Gundlach AL. Localization of relaxin-3 in brain of Macaca
fascicularis: identification of a nucleus incertus in primate. J Comp Neurol 517: 856–
872, 2009.
333. Ma S, Shen PJ, Burazin TC, Tregear GW, Gundlach AL. Comparative localization of
leucine-rich repeat-containing G-protein-coupled receptor-7 (RXFP1) mRNA and
[
33
P]-relaxin binding sites in rat brain: restricted somatic co-expression a clue to
relaxin action? Neuroscience 141: 329 –344, 2006.
334. Ma S, Shen PJ, Sang Q, Lanciego JL, Gundlach AL. Distribution of relaxin-3 mRNA and
immunoreactivity and RXFP3-binding sites in the brain of the macaque, Macaca fas-
cicularis.Ann NY Acad Sci 1160: 256 –258, 2009.
335. MacLennan AH. Pelvic girdle relaxation, developmental dysplasia of the hip and the
hormone relaxin–are they loosely connected?. In: Relaxin 2000: Proceedings of the
Third International Conference on Relaxin and Related Peptides , edited by Tregear GW,
Ivell R, Bathgate RA, Wade JD. Dordrecht, Germany: Kluwer, 2000, p. 407– 413.
336. MacLennan AH, Grant P, Borthwick AC. Relaxin and relaxin c-peptide levels in human
reproductive tissues. Reprod Fertil Dev 3: 577–583, 1991.
337. MacLennan AH, Grant P, Bryant-Greenwood G. hRLX-1 In vitro response of human
and pig myometrium. J Reprod Med 40: 703–706, 1995.
338. MacLennan AH, Grant P, Ness D, Down A. Effect of porcine relaxin and progesterone
on rat, pig and human myometrial activity in vitro. J Reprod Med 31: 43– 49, 1986.
339. MacLennan AH, Green RC, Grant P, Nicolson R. Ripening of the human cervix and
induction of labor with intracervical purified porcine relaxin. Obstet Gynecol 68: 598 –
601, 1986.
340. MacLennan AH, Nicolson R, Green RC, Bath M. Serum relaxin and pelvic pain of
pregnancy. Lancet 2: 243–245, 1986.
341. Martiny-Baron G, Kazanietz MG, Mischak H, Blumberg PM, Kochs G, Hug H, Marme
D, Schachtele C. Selective inhibition of protein kinase C isozymes by the indolocar-
bazole Go 6976. J Biol Chem 268: 9194 –9197, 1993.
342. Maseelall PB, Seungdamrong A, Weiss G, Wojtczuk AS, Donnelly R, Stouffer RL,
Goldsmith LT. Expression of LGR7 in the primate corpus luteum implicates the
corpus luteum as a relaxin target organ. Ann NY Acad Sci 1160: 147–151, 2009.
343. Masini E, Bani D, Bello MG, Bigazzi M, Mannaioni PF, Sacchi TB. Relaxin counteracts
myocardial damage induced by ischemia-reperfusion in isolated guinea pig hearts:
evidence for an involvement of nitric oxide. Endocrinology 138: 4713– 4720, 1997.
344. Masini E, Bani D, Bigazzi M, Mannaioni PF, Bani-Sacchi T. Effects of relaxin on mast
cells In vitro and in vivo studies in rats and guinea pigs. J Clin Invest 94: 1974–1980,
1994.
345. Masini E, Nistri S, Vannacci A, Bani Sacchi T, Novelli A, Bani D. Relaxin inhibits the
activation of human neutrophils: involvement of the nitric oxide pathway. Endocrinol-
ogy 145: 1106 –1112, 2004.
346. Masini E, Zagli G, Ndisang JF, Solazzo M, Mannaioni PF, Bani D. Protective effect of
relaxin in cardiac anaphylaxis: involvement of the nitric oxide pathway. Br J Pharmacol
137: 337–344, 2002.
347. Massague J. TGF-beta signal transduction. Annu Rev Biochem 67: 753–791, 1998.
348. Massicotte G, Parent A, JSL. Blunted responses to vasoconstrictors in mesenteric
vasculature but not in portal vein of spontaneously hypertensive rats treated with
relaxin. Proc Soc Exp Biol Med 190: 254 –259, 1989.
349. Masterson R, Hewitson TD, Kelynack K, Martic M, Parry L, Bathgate R, Darby I,
Becker G. Relaxin down-regulates renal fibroblast function and promotes matrix
remodelling in vitro. Nephrol Dial Transplant 19: 544 –552, 2004.
350. Mathieu MN, Wade JD, Tregear GW, Bond CP, Summers RJ, Catimel B, Nice EC,
Otvos L. Synthesis, conformational studies and biological activity of Nualpha-mono-
biotinylated rat relaxin. J Pept Res 57: 374 –382, 2001.
351. Mathieu P, Rahier J, Thomas K. Localization of relaxin in human gestational corpus
luteum. Cell Tissue Res 219: 213–216, 1981.
352. Matsumoto M, Kamohara M, Sugimoto T, Hidaka K, Takasaki J, Saito T, Okada M,
Yamaguchi T, Furuichi K. The novel G-protein coupled receptor SALPR shares se-
quence similarity with somatostatin and angiotensin receptors. Gene 248: 183–189,
2000.
353. Maynard SE, Min JY, Merchan J, Lim KH, Li J, Mondal S, Libermann TA, Morgan JP,
Sellke FW, Stillman IE, Epstein FH, Sukhatme VP, Karumanchi SA. Excess placental
soluble fms-like tyrosine kinase 1 (sFlt1) may contribute to endothelial dysfunction,
hypertension, and proteinuria in preeclampsia. J Clin Invest 111: 649 – 658, 2003.
354. Mazella J, Tang M, Tseng L. Disparate effects of relaxin and TGFbeta1: relaxin in-
creases, but TGFbeta1 inhibits, the relaxin receptor and the production of IGFBP-1 in
human endometrial stromal/decidual cells. Hum Reprod 19: 1513–1518, 2004.
355. Mazoujian G, Bryant GG. Relaxin in breast tissue. Lancet 335: 298 –299, 1990.
356. McDonald GA, Sarkar P, Rennke H, Unemori E, Kalluri R, Sukhatme VP. Relaxin
increases ubiquitin-dependent degradation of fibronectin in vitro and ameliorates
renal fibrosis in vivo. Am J Physiol Renal Physiol 285: F59 –F67, 2003.
357. McGowan BM, Stanley SA, Smith KL, Minnion JS, Donovan J, Thompson EL, Patterson
M, Connolly MM, Abbott CR, Small CJ, Gardiner JV, Ghatei MA, Bloom SR. Effects of
acute and chronic relaxin-3 on food intake and energy expenditure in rats. Regul Pept
136: 72–77, 2006.
358. McGowan BM, Stanley SA, Smith KL, White NE, Connolly MM, Thompson EL, Gar-
diner JV, Murphy KG, Ghatei MA, Bloom SR. Central relaxin-3 administration causes
hyperphagia in male Wistar rats. Endocrinology 146: 3295–3300, 2005.
359. McGuane JT, Danielson LA, Debrah JE, Rubin JP, Novak J, Conrad KP. Angiogenic
growth factors are new and essential players in the sustained relaxin vasodilatory
pathway in rodents and humans. Hypertension 57: 1151–1160, 2011.
360. McGuane JT, Debrah JE, Sautina L, Jarajapu YP, Novak J, Rubin JP, Grant MB, Segal M,
Conrad KP. Relaxin induces rapid dilation of rodent small renal and human subcuta-
neous arteries via PI3 kinase and nitric oxide. Endocrinology 152: 2786 –2796, 2011.
361. McKinley MJ, Allen AM, Burns P, Colvill LM, Oldfield BJ. Interaction of circulating
hormones with the brain: the roles of the subfornical organ and the organum vascu-
losum of the lamina terminalis. Clin Exp Pharmacol Physiol Suppl 25: S61– 67, 1998.
362. McKinley MJ, Burns P, Colvill LM, Oldfield BJ, Wade JD, Weisinger RS, Tregear GW.
Distribution of Fos immunoreactivity in the lamina terminalis and hypothalamus in-
duced by centrally administered relaxin in conscious rats. J Neuroendocrinol 9: 431–
437, 1997.
363. McMurtry JP, Kwok SC, Bryant-Greenwood GD. Target tissues for relaxin identified
in vitro with
125
I-labeled porcine relaxin. J Reprod Fertil 53: 209 –216, 1978.
364. Meera P, Anwer K, Monga M, Oberti C, Stefani E, Toro L, Sanborn BM. Relaxin
stimulates myometrial calcium-activated potassium channel activity via protein kinase
A. Am J Physiol Cell Physiol 269: C312–C317, 1995.
365. Min G, Hartzog MG, Jennings RL, Winn RJ, Sherwood OD. Evidence that endogenous
relaxin promotes growth of the vagina and uterus during pregnancy in gilts. Endocri-
nology 138: 560 –565, 1997.
366. Misrahi M, Beau I, Ghinea N, Vannier B, Loosfelt H, Meduri G, Vu Hai MT, Milgrom E.
The LH/CG and FSH receptors: different molecular forms and intracellular traffic. Mol
Cell Endocrinol 125: 161–167, 1996.
367. Miyamoto Y, Watanabe Y, Tanaka M. Developmental expression and serotonergic
regulation of relaxin 3/INSL7 in the nucleus incertus of rat brain. Regul Pept 145:
54 –59, 2008.
368. Mock P, Frydman R, Bellet D, Chassin D, Bischof P, Campana A, Bidart JM. Expression
of pro-EPIL peptides encoded by the insulin-like 4 (INSL4) gene in chromosomally
abnormal pregnancies. J Clin Endocrinol Metab 85: 3941–3944, 2000.
369. Mock P, Frydman R, Bellet D, Diawara DA, Lavaissiere L, Troalen F, Bidart JM.
Pro-EPIL forms are present in amniotic fluid and maternal serum during normal
pregnancy. J Clin Endocrinol Metab 84: 2253–2256, 1999.
BATHGATE ET AL.
474 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
370. Mookerjee I, Hewitson TD, Halls ML, Summers RJ, Mathai ML, Bathgate RA, Tregear
GW, Samuel CS. Relaxin inhibits renal myofibroblast differentiation via RXFP1, the
nitric oxide pathway, and Smad2. FASEB J 23: 1219 –1229, 2009.
371. Mookerjee I, Solly NR, Royce SG, Tregear GW, Samuel CS, Tang ML. Endogenous
relaxin regulates collagen deposition in an animal model of allergic airway disease.
Endocrinology 147: 754 –761, 2006.
372. Morelli SS, Petraglia F, Weiss G, Luisi S, Florio P, Wojtczuk A, Goldsmith LT. Endo-
metrial expression of relaxin and relaxin receptor in endometriosis. Fertil Steril 94:
2885–2887, 2010.
373. Morest DK. Connexions of the dorsal tegmental nucleus in rat and rabbit. J Anat 95:
229 –246, 1961.
374. Morikawa Y, Ueyama E, Senba E. Fasting-induced activation of mitogen-activated
protein kinases (ERK/p38) in the mouse hypothalamus. J Neuroendocrinol 16: 105–
112, 2004.
375. Muda M, He C, Martini PG, Ferraro T, Layfield S, Taylor D, Chevrier C, Schweick-
hardt R, Kelton C, Ryan PL, Bathgate RA. Splice variants of the relaxin and INSL3
receptors reveal unanticipated molecular complexity. Mol Hum Reprod 11: 591– 600,
2005.
376. Munro J, Skrobot O, Sanyoura M, Kay V, Susce MT, Glaser PE, de Leon J, Blakemore
AI, Arranz MJ. Relaxin polymorphisms associated with metabolic disturbance in pa-
tients treated with antipsychotics. J Psychopharmacol 26: 374 –379, 2012.
377. Nakamura K, Yamashita S, Omori Y, Minegishi T. A splice variant of the human
luteinizing hormone (LH) receptor modulates the expression of wild-type human LH
receptor. Mol Endocrinol 18: 1461–1470, 2004.
378. Nakanishi H, Brewer KA, Exton JH. Activation of the zeta isozyme of protein kinase C
by phosphatidylinositol 3,4,5-trisphosphate. J Biol Chem 268: 13–16, 1993.
379. Nef S, Parada LF. Cryptorchidism in mice mutant for INSL3. Nat Genet 22: 295–299,
1999.
380. Nef S, Shipman T, Parada LF. A molecular basis for estrogen-induced cryptorchidism.
Dev Biol 224: 354 –361, 2000.
381. Nguyen BT, Dessauer CW. Relaxin stimulates cAMP production in MCF-7 cells upon
overexpression of type V adenylyl cyclase. Ann NY Acad Sci 1041: 296 –299, 2005.
382. Nguyen BT, Dessauer CW. Relaxin stimulates protein kinase C
translocation: re-
quirement for cyclic adenosine 3=,5=-monophosphate production. Mol Endocrinol 19:
1012–1023, 2005.
383. Nguyen BT, Yang L, Sanborn BM, Dessauer CW. Phosphoinositide 3-kinase activity is
required for biphasic stimulation of cyclic adenosine 3=,5=-monophosphate by relaxin.
Mol Endocrinol 17: 1075–1084, 2003.
384. Nguyen MT, Showalter PR, Timmons CF, Nef S, Parada LF, Baker LA. Effects of
orchiopexy on congenitally cryptorchid insulin-3 knockout mice. J Urol 168: 1779 –
1783, 2002.
385. Nistri S, Bani D. Relaxin receptors and nitric oxide synthases: search for the missing
link. Reprod Biol Endocrinol 1: 5, 2003.
386. Nistri S, Chiappini L, Sassoli C, Bani D. Relaxin inhibits lipopolysaccharide-induced
adhesion of neutrophils to coronary endothelial cells by a nitric oxide-mediated mech-
anism. FASEB J 17: 2109 –2111, 2003.
387. Novak J, Danielson LA, Kerchner LJ, Sherwood OD, Ramirez RJ, Moalli PA, Conrad
KP. Relaxin is essential for renal vasodilation during pregnancy in conscious rats. J Clin
Invest 107: 1469 –1475, 2001.
388. Novak J, Parry LJ, Matthews JE, Kerchner LJ, Indovina K, Hanley-Yanez K, Doty KD,
Debrah DO, Shroff SG, Conrad KP. Evidence for local relaxin ligand-receptor expres-
sion and function in arteries. FASEB J 20: 2352–2362, 2006.
389. Novak J, Ramirez RJ, Gandley RE, Sherwood OD, Conrad KP. Myogenic reactivity is
reduced in small renal arteries isolated from relaxin-treated rats. Am J Physiol Regul
Integr Comp Physiol 283: R349 –R355, 2002.
390. Nunez A, Cervera-Ferri A, Olucha-Bordonau F, Ruiz-Torner A, Teruel V. Nucleus
incertus contribution to hippocampal theta rhythm generation. Eur J Neurosci 23:
2731–2738, 2006.
391. O’Byrne EM, Carriere BT, Sorensen L, Segaloff A, Schwabe C, Steinetz BG. Plasma
immunoreactive relaxin levels in pregnant and nonpregnant women. J Clin Endocrinol
Metab 47: 1106 –1110, 1978.
392. O’Byrne EM, Flitcraft JF, Sawyer WK, Hochman J, Weiss G, Steinetz BG. Relaxin
bioactivity and immunoactivity in human corpora lutea. Endocrinology 102: 1641–
1644, 1978.
393. O’Byrne EM, Steinetz BG. Radioimmunoassay (RIA) of relaxin in sera of various
species using an antiserum to porcine relaxin. Proc Soc Exp Biol Med 152: 272–276,
1976.
394. O’Connor WB, Cain GD, Zarrow MX. Elongation of the interpubic ligament in the
little brown bat (Myotis lucifugus). Proc Soc Exp Biol Med 123: 935–937, 1966.
395. Olucha-Bordonau FE, Teruel V, Barcia-Gonzalez J, Ruiz-Torner A, Valverde-Navarro
AA, Martinez-Soriano F. Cytoarchitecture and efferent projections of the nucleus
incertus of the rat. J Comp Neurol 464: 62–97, 2003.
396. Omi EC, Zhao S, Shanks RD, Sherwood OD. Evidence that systemic relaxin promotes
moderate water consumption during late pregnancy in rats. J Endocrinol 153: 33– 40,
1997.
397. Ono Y, Fujii T, Ogita K, Kikkawa U, Igarashi K, Nishizuka Y. Protein kinase C zeta
subspecies from rat brain: its structure, expression, and properties. Proc Natl Acad Sci
USA 86: 3099 –3103, 1989.
398. Orencole SF, Dinarello CA. Characterization of a subclone (D10S) of the D10. G41
helper T-cell line which proliferates to attomolar concentrations of interleukin-1 in
the absence of mitogens. Cytokine 1: 14 –22, 1989.
399. Osa T, Inoue H, Okabe K. Effects of porcine relaxin on contraction, membrane
response and cyclic AMP content in rat myometrium in comparison with the effects of
isoprenaline and forskolin. Br J Pharmacol 104: 950 –960, 1991.
400. Osheroff PL. Relaxin receptors in the rat uterus, brain, and heart. In: Progress in Relaxin
Research: Proceedings of the Second International Congress on the Homone Relaxin ,
edited by MacLennan AH, Tregear GW and Bryant-Greenwood GD. Adelaide, Aus-
tralia: Global Publications, 1995, p. 159 –176.
401. Osheroff PL, Cronin MJ, Lofgren JA. Relaxin binding in the rat heart atrium. Proc Natl
Acad Sci USA 89: 2384 –2388, 1992.
402. Osheroff PL, Ho WH. Expression of relaxin mRNA and relaxin receptors in postnatal
and adult rat brains and hearts. Localization and developmental patterns. J Biol Chem
268: 15193–15199, 1993.
403. Osheroff PL, Ling VT, Vandlen RL, Cronin MJ, Lofgren JA. Preparation of biologically
active 32P-labeled human relaxin. Displaceable binding to rat uterus, cervix, and
brain. J Biol Chem 265: 9396 –9401, 1990.
404. Osheroff PL, Phillips HS. Autoradiographic localization of relaxin binding sites in rat
brain. Proc Natl Acad Sci USA 88: 6413– 6417, 1991.
405. Otsubo H, Onaka T, Suzuki H, Katoh A, Ohbuchi T, Todoroki M, Kobayashi M,
Fujihara H, Yokoyama T, Matsumoto T, Ueta Y. Centrally administered relaxin-3
induces Fos expression in the osmosensitive areas in rat brain and facilitates water
intake. Peptides 31: 1124 –1130, 2010.
406. Overbeek PA, Gorlov IP, Sutherland RW, Houston JB, Harrison WR, Boettger-Tong
HL, Bishop CE, Agoulnik AI. A transgenic insertion causing cryptorchidism in mice.
Genesis 30: 26 –35, 2001.
407. Palejwala S, Stein D, Wojtczuk A, Weiss G, Goldsmith LT. Demonstration of a relaxin
receptor and relaxin-stimulated tyrosine phosphorylation in human lower uterine
segment fibroblasts. Endocrinology 139: 1208 –1212, 1998.
408. Palejwala S, Stein DE, Weiss G, Monia BP, Tortoriello D, Goldsmith LT. Relaxin
positively regulates matrix metalloproteinase expression in human lower uterine seg-
ment fibroblasts using a tyrosine kinase signaling pathway. Endocrinology 142: 3405–
3413, 2001.
409. Palejwala S, Tseng L, Wojtczuk A, Weiss G, Goldsmith LT. Relaxin gene and protein
expression and its regulation of procollagenase and vascular endothelial growth factor
in human endometrial cells. Biol Reprod 66: 1743–1748, 2002.
410. Park JI, Semyonov J, Yi W, Chang CL, Hsu SY. Regulation of receptor signaling by
relaxin A chain motifs: derivation of pan-specific and LGR7-specific human relaxin
analogs. J Biol Chem 283: 32099 –32109, 2008.
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
475Physiol Rev VOL 93 JANUARY 2013 www.prv.org
411. Parsell DA, Mak JY, Amento EP, Unemori EN. Relaxin binds to and elicits a response
from cells of the human monocytic cell line, THP-1. J Biol Chem 271: 27936 –27941,
1996.
412. Paust HJ, Wessels J, Ivell R, Mukhopadhyay AK. The expression of the RLF/INSL3 gene
is reduced in Leydig cells of the aging rat testis. Exp Gerontol 37: 1461–1467, 2002.
413. Paxinos G, Watson C. The Rat Brain in Stereotaxic Coordinates. New York: Academic,
1998.
414. Pepe A, Ferlin A, Gianesello L, Facciolli A, Agoulnik AI, Foresta C. INSL3 plays a role
in the balance between bone formation and resorption. Ann NY Acad Sci 1160: 219 –
220, 2009.
415. Petersen LK, Svane D, Uldbjerg N, Forman A. Effects of human relaxin on isolated rat
and human myometrium and uteroplacental arteries. Obstet Gynecol 78: 757–762,
1991.
416. Piccenna L, Shen PJ, Ma S, Burazin TC, Gossen JA, Mosselman S, Bathgate RA, Gund-
lach AL. Localization of LGR7 gene expression in adult mouse brain using LGR7
knock-out/LacZ knock-in mice: correlation with LGR7 mRNA distribution. Ann NY
Acad Sci 1041: 197–204, 2005.
417. Piedras-Renteria ES, Sherwood OD, Best PM. Effects of relaxin on rat atrial myocytes.
I. Inhibition of I(to) via PKA-dependent phosphorylation. Am J Physiol Heart Circ Physiol
272: H1791–H1797, 1997.
418. Piedras-Renteria ES, Sherwood OD, Best PM. Effects of relaxin on rat atrial myocytes.
II. Increased calcium influx derived from action potential prolongation. Am J Physiol
Heart Circ Physiol 272: H1798 –H1803, 1997.
419. Pietila EM, Tuusa JT, Apaja PM, Aatsinki JT, Hakalahti AE, Rajaniemi HJ, Petaja-Repo
UE. Inefficient maturation of the rat luteinizing hormone receptor: a putative way to
regulate receptor numbers at the cell surface. J Biol Chem 280: 26622–26629, 2005.
420. Pini A, Shemesh R, Samuel CS, Bathgate RA, Zauberman A, Hermesh C, Wool A, Bani
D, Rotman G. Prevention of bleomycin-induced pulmonary fibrosis by a novel antifi-
brotic peptide with relaxin-like activity. J Pharmacol Exp Ther 335: 589 –599, 2010.
421. Ponikowski P, Metra M, Teerlink JR, Unemori E, Felker GM, Voors AA, Filippatos G,
Greenberg B, Teichman SL, Severin T, Mueller-Velten G, Cotter G, Davison BA.
Design of the RELAXin in acute heart failure study. Am Heart J 163: 149 –155, 2012.
422. Potter E, Sutton S, Donaldson C, Chen R, Perrin M, Lewis K, Sawchenko PE, Vale W.
Distribution of corticotropin-releasing factor receptor mRNA expression in the rat
brain and pituitary. Proc Natl Acad Sci USA 91: 8777– 8781, 1994.
423. Price MA, Cruzalegui FH, Treisman R. The p38 and ERK MAP kinase pathways
cooperate to activate ternary complex factors and c-fos transcription in response to
UV light. EMBO J 15: 6552– 6563, 1996.
424. Pusch W, Balvers M, Ivell R. Molecular cloning and expression of the relaxin-like factor
from the mouse testis. Endocrinology 137: 3009 –3013, 1996.
425. Quaggin SE, Schwartz L, Cui S, Igarashi P, Deimling J, Post M, Rossant J. The basic-
helix-loop-helix protein pod1 is critically important for kidney and lung organogenesis.
Development 126: 5771–5783, 1999.
426. Quagliarello J, Goldsmith L, Steinetz B, Lustig DS, Weiss G. Induction of relaxin
secretion in nonpregnant women by human chorionic gonadotropin. J Clin Endocrinol
Metab 51: 74 –77, 1980.
427. Quattrone S, Chiappini L, Scapagnini G, Bigazzi B, Bani D. Relaxin potentiates the
expression of inducible nitric oxide synthase by endothelial cells from human umbilical
vein in in vitro culture. Mol Hum Reprod 10: 325–330, 2004.
428. Quintana J, Hipkin RW, Ascoli M. A polyclonal antibody to a synthetic peptide derived
from the rat follicle-stimulating hormone receptor reveals the recombinant receptor
as a 74-kilodalton protein. Endocrinology 133: 2098 –2104, 1993.
429. Radestock Y, Willing C, Kehlen A, Hoang-Vu C, Hombach-Klonisch S. Relaxin en-
hances S100A4 and promotes growth of human thyroid carcinoma cell xenografts.
Mol Cancer Res 8: 494 –506, 2010.
430. Rands E, Candelore MR, Cheung AH, Hill WS, Strader CD, Dixon RA. Mutational
analysis of beta-adrenergic receptor glycosylation. J Biol Chem 265: 10759 –10764,
1990.
431. Ren Q, Kurose H, Lefkowitz RJ, Cotecchia S. Constitutively active mutants of the
alpha 2-adrenergic receptor. J Biol Chem 268: 16483–16487, 1993.
432. Roche PJ, Butkus A, Wintour EM, Tregear G. Structure and expression of Leydig
insulin-like peptide mRNA in the sheep. Mol Cell Endocrinol 121: 171–177, 1996.
433. Rosengren KJ, Bathgate RA, Craik DJ, Daly NL, Haugaard-Jonsson LM, Hossain MA,
Wade JD. Structural insights into the function of relaxins. Ann NY Acad Sci 1160:
20 –26, 2009.
434. Rosengren KJ, Lin F, Bathgate RA, Tregear GW, Daly NL, Wade JD, Craik DJ. Solution
structure and novel insights into the determinants of the receptor specificity of human
relaxin-3. J Biol Chem 281: 5845–5851, 2006.
435. Rosengren KJ, Zhang S, Lin F, Daly NL, Scott DJ, Hughes RA, Bathgate RA, Craik DJ,
Wade JD. Solution structure and characterization of the LGR8 receptor binding sur-
face of insulin-like peptide 3. J Biol Chem 281: 28287–28295, 2006.
436. Roux PP, Blenis J. ERK and p38 MAPK-activated protein kinases: a family of protein
kinases with diverse biological functions. Microbiol Mol Biol Rev 68: 320 –344, 2004.
437. Royce SG, Miao YR, Lee M, Samuel CS, Tregear GW, Tang ML. Relaxin reverses
airway remodeling and airway dysfunction in allergic airways disease. Endocrinology
150: 2692–2699, 2009.
438. Sadana R, Dessauer CW. Physiological roles for G protein-regulated adenylyl cyclase
isoforms: insights from knockout and overexpression studies. Neurosignals 17: 5–22,
2009.
439. Sakbun V, Ali SM, Greenwood FC, Bryant GG. Human relaxin in the amnion, chorion,
decidua parietalis, basal plate, and placental trophoblast by immunocytochemistry and
northern analysis. J Clin Endocrinol Metab 70: 508 –514, 1990.
440. Samuel CS, Hewitson TD, Zhang Y, Kelly DJ. Relaxin ameliorates fibrosis in experi-
mental diabetic cardiomyopathy. Endocrinology 149: 3286 –3293, 2008.
441. Samuel CS, Royce SG, Chen B, Cao H, Gossen JA, Tregear GW, Tang ML. Relaxin
family peptide receptor-1 protects against airway fibrosis during homeostasis but not
against fibrosis associated with chronic allergic airways disease. Endocrinology 150:
1495–1502, 2009.
442. Samuel CS, Tian H, Zhao L, Amento EP. Relaxin is a key mediator of prostate growth
and male reproductive tract development. Lab Invest 83: 1055–1067, 2003.
443. Samuel CS, Unemori EN, Mookerjee I, Bathgate RA, Layfield SL, Mak J, Tregear GW,
Du XJ. Relaxin modulates cardiac fibroblast proliferation, differentiation, and collagen
production and reverses cardiac fibrosis in vivo. Endocrinology 145: 4125– 4133, 2004.
444. Samuel CS, Zhao C, Bathgate RA, Bond CP, Burton MD, Parry LJ, Summers RJ, Tang
ML, Amento EP, Tregear GW. Relaxin deficiency in mice is associated with an age-
related progression of pulmonary fibrosis. FASEB J 17: 121–123, 2003.
445. Samuel CS, Zhao C, Bond CP, Hewitson TD, Amento EP, Summers RJ. Relaxin-1-
deficient mice develop an age-related progression of renal fibrosis. Kidney Int 65:
2054 –2064, 2004.
446. Sanborn BM, Kuo HS, Weisbrodt NW, Sherwood OD. The interaction of relaxin with
the rat uterus. I. Effect on cyclic nucleotide levels and spontaneous contractile activity.
Endocrinology 106: 1210 –1215, 1980.
447. Sanchez-Blazquez P, Garzon J. alpha N-acetyl-beta-endorphin-(1–31) disrupts the
diminishing effect of mastoparan on opioid- and clonidine-evoked supraspinal antino-
ciception in mice. J Pharmacol Exp Ther 273: 787–792, 1995.
448. Sasaguri K, Kikuchi M, Hori N, Yuyama N, Onozuka M, Sato S. Suppression of stress
immobilization-induced phosphorylation of ERK 1/2 by biting in the rat hypothalamic
paraventricular nucleus. Neurosci Lett 383: 160 –164, 2005.
449. Sato M, Horinouchi T, Hutchinson DS, Evans BA, Summers RJ. Ligand-directed sig-
naling at the
3
-adrenoceptor produced by 3-(2-Ethylphenoxy)-1-[(1,S)-1,2,3,4-tet-
rahydronapth-1-ylamino]-2S-2-propan ol oxalate (SR59230A) relative to receptor
agonists. Mol Pharmacol 72: 1359 –1368, 2007.
450. Schink W, Struck H. Relaxin in the Allen-Doisy test. Zentralblatt fur Gynakologie 90:
675– 678, 1968.
451. Schmidt CL, Black VH, Sarosi P, Weiss G. Progesterone and relaxin secretion in
relation to the ultrastructure of human luteal cells in culture: effects of human chori-
onic gonadotropin. Am J Obstet Gynecol 155: 1209 –1219, 1986.
452. Schwabe C, Bullesbach EE. Relaxin: structures, functions, promises, nonevolution.
FASEB J 8: 1152–1160, 1994.
BATHGATE ET AL.
476 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
453. Scott DJ, Fu P, Shen PJ, Gundlach A, Layfield S, Riesewijk A, Tomiyama H, Hutson JM,
Tregear GW, Bathgate RA. Characterization of the rat INSL3 receptor. Ann NY Acad
Sci 1041: 13–16, 2005.
454. Scott DJ, Layfield S, Riesewijk A, Morita H, Tregear GW, Bathgate RA. Characteriza-
tion of the mouse and rat relaxin receptors. Ann NY Acad Sci 1041: 8 –12, 2005.
455. Scott DJ, Layfield S, Riesewijk A, Morita H, Tregear GW, Bathgate RA. Identification
and characterization of the mouse and rat relaxin receptors as the novel orthologues
of human leucine-rich repeat-containing G-protein-coupled receptor 7. Clin Exp Phar-
macol Physiol 31: 828 – 832, 2004.
456. Scott DJ, Layfield S, Yan Y, Sudo S, Hsueh AJ, Tregear GW, Bathgate RA. Character-
ization of novel splice variants of LGR7 and LGR8 reveals that receptor signaling is
mediated by their unique LDLa modules. J Biol Chem 281: 34942–34954, 2006.
457. Scott DJ, Tregear GW, Bathgate RA. LGR7-truncate is a splice variant of the relaxin
receptor LGR7 and is a relaxin antagonist in vitro. Ann NY Acad Sci 1041: 22–26, 2005.
458. Scott DJ, Tregear GW, Bathgate RA. Modeling the primary hormone-binding site of
RXFP1 and RXFP2. Ann NY Acad Sci 1160: 74 –77, 2009.
459. Scott DJ, Wilkinson TN, Zhang S, Ferraro T, Wade JD, Tregear GW, Bathgate RA.
Defining the LGR8 residues involved in binding insulin-like peptide 3. Mol Endocrinol
21: 1699 –1712, 2007.
460. Sedaghat K, Shen PJ, Finkelstein DI, Henderson JM, Gundlach AL. Leucine-rich re-
peat-containing G-protein-coupled receptor 8 in the rat brain: enrichment in thalamic
neurons and their efferent projections. Neuroscience 156: 319 –333, 2008.
461. Seibold JR. Relaxins: lessons and limitations. Curr Rheumatol Rep 4: 275–276, 2002.
462. Seibold JR, Korn JH, Simms R, Clements PJ, Moreland LW, Mayes MD, Furst DE,
Rothfield N, Steen V, Weisman M, Collier D, Wigley FM, Merkel PA, Csuka ME, Hsu
V, Rocco S, Erikson M, Hannigan J, Harkonen WS, Sanders ME. Recombinant human
relaxin in the treatment of scleroderma. A randomized, double-blind, placebo-con-
trolled trial. Ann Intern Med 132: 871– 879, 2000.
463. Shabanpoor F, Akhter Hossain M, Ryan PJ, Belgi A, Layfield S, Kocan M, Zhang S,
Samuel CS, Gundlach AL, Bathgate RA, Separovic F, Wade JD. Minimization of
human relaxin-3 leading to high-affinity analogues with increased selectivity for
relaxin-family peptide 3 receptor (RXFP3) over RXFP1. J Med Chem 55: 1671–
1681, 2012.
464. Shabanpoor F, Bathgate RA, Hossain MA, Giannakis E, Wade JD, Hughes RA. Design,
synthesis and pharmacological evaluation of cyclic mimetics of the insulin-like peptide
3 (INSL3) B-chain. J Pept Sci 13: 113–120, 2007.
465. Shah AM, Mann DL. In search of new therapeutic targets and strategies for heart
failure: recent advances in basic science. Lancet 378: 704 –712, 2011.
466. Shemesh R, Hermesh C, Toporik A, Levine Z, Novik A, Wool A, Kliger Y, Rosenberg
A, Bathgate RA, Cohen Y. Activation of relaxin-related receptors by short, linear
peptides derived from a collagen-containing precursor. Ann NY Acad Sci 1160: 78 – 86,
2009.
467. Shemesh R, Toporik A, Levine Z, Hecht I, Rotman G, Wool A, Dahary D, Gofer E,
Kliger Y, Soffer MA, Rosenberg A, Eshel D, Cohen Y. Discovery and validation of novel
peptide agonists for G-protein-coupled receptors. J Biol Chem 283: 34643–34649,
2008.
468. Shen CP, Tsimberg Y, Salvadore C, Meller E. Activation of Erk and JNK MAPK
pathways by acute swim stress in rat brain regions. BMC Neurosci 5: 36, 2004.
469. Shen PJ, Fu P, Phelan KD, Scott DJ, Layfield S, Tregear GW, Bathgate RA, Gundlach
AL. Restricted expression of LGR8 in intralaminar thalamic nuclei of rat brain suggests
a role in sensorimotor systems. Ann NY Acad Sci 1041: 510 –515, 2005.
470. Sherwood OD. Relaxin. In: The Physiology of Reproduction , edited by Knobil E, Neill JD.
New York: Raven, 1994, p. 861–1009.
471. Sherwood OD. Relaxin’s physiological roles and other diverse actions. Endocr Rev 25:
205–234, 2004.
472. Sherwood OD, Crnekovic VE, Gordon WL, Rutherford JE. Radioimmunoassay of
relaxin throughout pregnancy and during parturition in the rat. Endocrinology 107:
691– 698, 1980.
473. Sherwood OD, Rutherford JE. Relaxin immunoactivity levels in ovarian extracts ob-
tained from rats during various reproductive states and from adult cycling pigs. Endo-
crinology 108: 1171–1177, 1981.
474. Shirota K, Tateishi K, Koji T, Hishikawa Y, Hachisuga T, Kuroki M, Kawarabayashi T.
Early human preantral follicles have relaxin and relaxin receptor (LGR7), relaxin pro-
motes their development. J Clin Endocrinol Metab 90: 516 –521, 2005.
475. Shpakov AO, Gur’yanov IA, Kuznetsova LA, Plesneva SA, Shpakova EA, Vlasov GP,
Pertseva MN. Studies of the molecular mechanisms of action of relaxin on the adenylyl
cyclase signaling system using synthetic peptides derived from the LGR7 relaxin re-
ceptor. Neurosci Behav Physiol 37: 705–714, 2007.
476. Shymko RM, De Meyts P, Thomas R. Logical analysis of timing-dependent receptor
signaling specificity: application to the insulin receptor metabolic and mitogenic sig-
naling pathways. Biochem J 326: 463– 469, 1997.
477. Siebel AL, Gehring HM, Reytomas IG, Parry LJ. Inhibition of oxytocin receptor and
estrogen receptor-alpha expression, but not relaxin receptors (LGR7), in the myo-
metrium of late pregnant relaxin gene knockout mice. Endocrinology 144: 4272– 4275,
2003.
478. Silvertown JD, Neschadim A, Liu HN, Shannon P, Walia JS, Kao JC, Robertson J,
Summerlee AJ, Medin JA. Relaxin-3 and receptors in the human and rhesus brain and
reproductive tissues. Regul Pept 2009.
479. Silvertown JD, Summerlee AJ, Klonisch T. Relaxin-like peptides in cancer. Int J Cancer
107: 513–519, 2003.
480. Silvertown JD, Symes JC, Neschadim A, Nonaka T, Kao JC, Summerlee AJ, Medin JA.
Analog of H2 relaxin exhibits antagonistic properties and impairs prostate tumor
growth. FASEB J 21: 754 –765, 2007.
481. Simon C, Einspanier A. The hormonal induction of cervical remodeling in the common
marmoset monkey (Callithrix jacchus). Reproduction 137: 517–525, 2009.
482. Singh S, Bennett RG. Relaxin signaling activates peroxisome proliferator-activated
receptor gamma. Mol Cell Endocrinol 315: 239 –245, 2010.
483. Sinnayah P, Burns P, Wade JD, Weisinger RS, McKinley MJ. Water drinking in rats
resulting from intravenous relaxin and its modification by other dipsogenic factors.
Endocrinology 140: 5082–5086, 1999.
484. Skroblin P, Grossmann S, Schafer G, Rosenthal W, Klussmann E. Mechanisms of
protein kinase A anchoring. Int Rev Cell Mol Biol 283: 235–330, 2010.
485. Smith CM, Lawrence AJ, Sutton SW, Gundlach AL. Behavioral phenotyping of mixed
background (129S5:B6) relaxin-3 knockout mice. Ann NY Acad Sci 1160: 236 –241,
2009.
486. Smith CM, Ryan PJ, Hosken IT, Ma S, Gundlach AL. Relaxin-3 systems in the brain-The
first 10 years. J Chem Neuroanat 2011.
487. Smith CM, Shen PJ, Banerjee A, Bonaventure P, Ma S, Bathgate RA, Sutton SW,
Gundlach AL. Distribution of relaxin-3 and RXFP3 within arousal, stress, affective, and
cognitive circuits of mouse brain. J Comp Neurol 518: 4016 – 4045, 2010.
488. Smith CM, Shen PJ, Ma S, Sutton SW, Gundlach AL. Verification of a relaxin-3 knock-
out/LacZ reporter mouse as a model of relaxin-3 deficiency. Ann NY Acad Sci 1160:
259 –260, 2009.
489. Sokol RZ, Wang XS, Lechago J, Johnston PD, Swerdloff RS. Immunohistochemical
localization of relaxin in human prostate. J Histochem Cytochem 37: 1253–1255, 1989.
490. Spanel-Borowski K, Schafer I, Zimmermann S, Engel W, Adham IM. Increase in final
stages of follicular atresia and premature decay of corpora lutea in INSL3-deficient
mice. Mol Reprod Dev 58: 281–286, 2001.
491. Spiess AN, Balvers M, Tena-Sempere M, Huhtaniemi I, Parry L, Ivell R. Structure and
expression of the rat relaxin-like factor (RLF) gene. Mol Reprod Dev 54: 319 –325,
1999.
492. St Louis J, Massicotte G. Chronic decrease of blood pressure by rat relaxin in spon-
taneously hypertensive rats. Life Sci 37: 1351–1357, 1985.
493. Standaert ML, Bandyopadhyay G, Kanoh Y, Sajan MP, Farese RV. Insulin and PIP3
activate PKC-
by mechanisms that are both dependent and independent of phos-
phorylation of activation loop (T410) and autophosphorylation (T560) sites. Biochem-
istry 40: 249 –255, 2001.
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
477Physiol Rev VOL 93 JANUARY 2013 www.prv.org
494. Stanley KK, Kocher HP, Luzio JP, Jackson P, Tschopp J. The sequence and topology of
human complement component C9. EMBO J 4: 375–382, 1985.
495. Steinetz BG, Beach VL, Blye RP, Kroc RL. Changes in the composition of the rat uterus
following a single injection of relaxin. Endocrinology 61: 287–292, 1957.
496. Steinetz BG, O’Byrne EM, Butler MC, Hickman LB. Hormonal regulation of the
connective tissue of the symphysis pubis. In: Biology of Relaxin and Its Role in the Human
, edited by Bigazzi M, Greenwood FC, Gasparri F. Amsterdam: Excerpta Medica,
1983, p. 71–92.
497. Stewart DR, Celniker AC, Taylor CA Jr, Cragun JR, Overstreet JW, Lasley BL. Relaxin
in the peri-implantation period. J Clin Endocrinol Metab 70: 1771–1773, 1990.
498. Stewart DR, Overstreet JW, Celniker AC, Hess DL, Cragun JR, Boyers SP, Lasley BL.
The relationship between hCG and relaxin secretion in normal pregnancies vs peri-
implantation spontaneous abortions. Clin Endocrinol 38: 379 –385, 1993.
499. Stewart DR, VandeVoort CA. Relaxin secretion by human granulosa cell culture is
predictive of in-vitro fertilization-embryo transfer success. Hum Reprod 14: 338 –344,
1999.
500. Stohr H, Berger C, Frohlich S, Weber BH. A novel gene encoding a putative trans-
membrane protein with two extracellular CUB domains and a low-density lipoprotein
class A module: isolation of alternatively spliced isoforms in retina and brain. Gene 286:
223–231, 2002.
501. Sudo S, Kumagai J, Nishi S, Layfield S, Ferraro T, Bathgate RA, Hsueh AJ. H3 relaxin is
a specific ligand for LGR7 and activates the receptor by interacting with both the
ectodomain and the exoloop 2. J Biol Chem 278: 7855–7862, 2003.
502. Summerlee AJ, Hornsby DJ, Ramsey DG. The dipsogenic effects of rat relaxin: the
effect of photoperiod and the potential role of relaxin on drinking in pregnancy.
Endocrinology 139: 2322–2328, 1998.
503. Summerlee AJ, Ramsey DG, Poterski RS. Neutralization of relaxin within the brain
affects the timing of birth in rats. Endocrinology 139: 479 – 484, 1998.
504. Sunn N, Egli M, Burazin TC, Burns P, Colvill L, Davern P, Denton DA, Oldfield BJ,
Weisinger RS, Rauch M, Schmid HA, McKinley MJ. Circulating relaxin acts on subfor-
nical organ neurons to stimulate water drinking in the rat. Proc Natl Acad Sci USA 99:
1701–1706, 2002.
505. Sunn N, McKinley MJ, Oldfield BJ. Identification of efferent neural pathways from the
lamina terminalis activated by blood-borne relaxin. J Neuroendocrinol 13: 432– 437,
2001.
506. Sutton RE, Koob GF, Le Moal M, Rivier J, Vale W. Corticotropin releasing factor
produces behavioural activation in rats. Nature 297: 331–333, 1982.
507. Sutton SW, Bonaventure P, Kuei C, Nepomuceno D, Wu J, Zhu J, Lovenberg TW, Liu
C. G-protein-coupled receptor (GPCR)-142 does not contribute to relaxin-3 binding
in the mouse brain: further support that relaxin-3 is the physiological ligand for
GPCR135. Neuroendocrinology 82: 139 –150, 2005.
508. Sutton SW, Bonaventure P, Kuei C, Roland B, Chen J, Nepomuceno D, Lovenberg
TW, Liu C. Distribution of G-protein-coupled receptor (GPCR)135 binding sites and
receptor mRNA in the rat brain suggests a role for relaxin-3 in neuroendocrine and
sensory processing. Neuroendocrinology 80: 298 –307, 2004.
509. Sutton SW, Shelton J, Smith C, Williams J, Yun S, Motley T, Kuei C, Bonaventure P,
Gundlach A, Liu C, Lovenberg T. Metabolic and neuroendocrine responses to RXFP3
modulation in the central nervous system. Ann NY Acad Sci 1160: 242–249, 2009.
510. Svendsen AM, Vrecl M, Ellis TM, Heding A, Kristensen JB, Wade JD, Bathgate RA, De
Meyts P, Nohr J. Cooperative binding of insulin-like Peptide 3 to a dimeric relaxin
family peptide receptor 2. Endocrinology 149: 1113–1120, 2008.
511. Svendsen AM, Zalesko A, Konig J, Vrecl M, Heding A, Kristensen JB, Wade JD, Bath-
gate RA, De Meyts P, Nohr J. Negative cooperativity in H2 relaxin binding to a dimeric
relaxin family peptide receptor 1. Mol Cell Endocrinol 296: 10 –17, 2008.
512. Tabanelli S, Tang B, Gurpide E. In vitro decidualization of human endometrial stromal
cells. J Steroid Biochem Mol Biol 42: 337–344, 1992.
513. Tan YY, Dawson NF, Kompa AR, Bond CP, Claasz A, Wade JD, Tregear GW, Sum-
mers RJ. Structural requirements for the interaction of sheep insulin-like factor 3 with
relaxin receptors in rat atria. Eur J Pharmacol 457: 153–160, 2002.
514. Tan YY, Wade JD, Tregear GW, Summers RJ. Comparison of relaxin receptors in rat
isolated atria and uterus by use of synthetic and native relaxin analogues. Br J Pharmacol
123: 762–770, 1998.
515. Tan YY, Wade JD, Tregear GW, Summers RJ. Quantitative autoradiographic studies of
relaxin binding in rat atria, uterus and cerebral cortex: characterization and effects of
oestrogen treatment. Br J Pharmacol 127: 91–98, 1999.
516. Tanaka M, Iijima N, Miyamoto Y, Fukusumi S, Itoh Y, Ozawa H, Ibata Y. Neurons
expressing relaxin 3/INSL 7 in the nucleus incertus respond to stress. Eur J Neurosci 21:
1659 –1670, 2005.
517. Tang B, Guller S, Gurpide E. Cyclic adenosine 3=,5=-monophosphate induces prolactin
expression in stromal cells isolated from human proliferative endometrium. Endocri-
nology 133: 2197–2203, 1993.
518. Tang ML, Samuel CS, Royce SG. Role of relaxin in regulation of fibrosis in the lung. Ann
NY Acad Sci 1160: 342–347, 2009.
519. Tang XM, Chegini N. Human fallopian tube as an extraovarian source of relaxin:
messenger ribonucleic acid expression and cellular localization of immunoreactive
protein and
125
I-relaxin binding sites. Biol Reprod 52: 1343–1349, 1995.
520. Tao YX, Johnson NB, Segaloff DL. Constitutive and agonist-dependent self-associa-
tion of the cell surface human lutropin receptor. J Biol Chem 279: 5904 –5914, 2004.
521. Tashima LS, Hieber AD, Greenwood FC, Bryant-Greenwood GD. The human Leydig
insulin-like (hLEY I-L) gene is expressed in the corpus luteum and trophoblast. J Clin
Endocrinol Metab 80: 707–710, 1995.
522. Tashima LS, Mazoujian G, Bryant-Greenwood GD. Human relaxins in normal, benign
and neoplastic breast tissue. J Mol Endocrinol 12: 351–364, 1994.
523. Tashima LS, Yamamoto SY, Yasuda M, Millar LK, Bryant-Greenwood GD. Decidual
relaxins: gene and protein up-regulation in preterm premature rupture of the mem-
branes by complementary DNA arrays and quantitative immunocytochemistry. Am J
Obstet Gynecol 187: 785–797, 2002.
524. Taylor MJ, Clark CL. Evidence for a novel source of relaxin: atrial cardiocytes. J
Endocrinol 143: R5–R8, 1994.
525. Teerds KJ, de Boer-Brouwer M, Dorrington JH, Balvers M, Ivell R. Identification of
markers for precursor and Leydig cell differentiation in the adult rat testis following
ethane dimethyl sulphonate administration. Biol Reprod 60: 1437–1445, 1999.
526. Teerlink JR, Metra M, Felker GM, Ponikowski P, Voors AA, Weatherley BD, Marmor
A, Katz A, Grzybowski J, Unemori E, Teichman SL, Cotter G. Relaxin for the treat-
ment of patients with acute heart failure (Pre-RELAX-AHF): a multicentre, ran-
domised, placebo-controlled, parallel-group, dose-finding phase IIb study. Lancet 373:
1429 –1439, 2009.
527. Teichman SL, Unemori E, Dschietzig T, Conrad K, Voors AA, Teerlink JR, Felker GM,
Metra M, Cotter G. Relaxin, a pleiotropic vasodilator for the treatment of heart
failure. Heart Fail Rev 14: 321–329, 2009.
528. Teichman SL, Unemori E, Teerlink JR, Cotter G, Metra M. Relaxin: review of biology
and potential role in treating heart failure. Curr Heart Fail Rep 7: 75– 82, 2010.
529. Telgmann R, Gellersen B. Marker genes of decidualization: activation of the decidual
prolactin gene. Hum Reprod Update 4: 472– 479, 1998.
530. Tena-Sempere M, Manna PR, Huhtaniemi I. Molecular cloning of the mouse follicle-
stimulating hormone receptor complementary deoxyribonucleic acid: functional ex-
pression of alternatively spliced variants and receptor inactivation by a C566T transi-
tion in exon 7 of the coding sequence. Biol Reprod 60: 1515–1527, 1999.
531. Thomas GR, Vandlen R. The purely chronotropic effects of relaxin in the rat isolated
heart. J Pharm Pharmacol 45: 927–928, 1993.
532. Thompson VC, Morris TG, Cochrane DR, Cavanagh J, Wafa LA, Hamilton T, Wang S,
Fazli L, Gleave ME, Nelson CC. Relaxin becomes upregulated during prostate cancer
progression to androgen independence and is negatively regulated by androgens.
Prostate 66: 1698 –1709, 2006.
533. Tomboc M, Lee PA, Mitwally MF, Schneck FX, Bellinger M, Witchel SF. Insulin-like
3/relaxin-like factor gene mutations are associated with cryptorchidism. J Clin Endo-
crinol Metab 85: 4013– 4018, 2000.
534. Toth M, Taskinen P, Ruskoaho H. Relaxin stimulates atrial natriuretic peptide secre-
tion in perfused rat heart. J Endocrinol 150: 487– 495, 1996.
BATHGATE ET AL.
478 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
535. Tsafriri A, Pomerantz SH. Oocyte maturation inhibitor. Clin Endocrinol Metab 15:
157–170, 1986.
536. Uberall F, Giselbrecht S, Hellbert K, Fresser F, Bauer B, Gschwendt M, Grunicke HH,
Baier G. Conventional PKC-
, novel PKC-and PKC-
, but not atypical PKC-
are
MARCKS kinases in intact NIH 3T3 fibroblasts. J Biol Chem 272: 4072– 4078, 1997.
537. Ueda H, Inoue M, Yoshida A, Mizuno K, Yamamoto H, Maruo J, Matsuno K, Mita S.
Metabotropic neurosteroid/sigma-receptor involved in stimulation of nociceptor
endings of mice. J Pharmacol Exp Ther 298: 703–710, 2001.
538. Unemori E, Sibai B, Teichman SL. Scientific rationale and design of a phase I safety study of
relaxin in women with severe preeclampsia. Ann NY Acad Sci 1160: 381–384, 2009.
539. Unemori EN, Amento EP. Relaxin modulates synthesis and secretion of procollage-
nase and collagen by human dermal fibroblasts. J Biol Chem 265: 10681–10685, 1990.
540. Unemori EN, Erikson ME, Rocco SE, Sutherland KM, Parsell DA, Mak J, Grove BH.
Relaxin stimulates expression of vascular endothelial growth factor in normal human
endometrial cells in vitro and is associated with menometrorrhagia in women. Hum
Reprod 14: 800 – 806, 1999.
541. Unemori EN, Lewis M, Constant J, Arnold G, Grove BH, Normand J, Deshpande U,
Salles A, Pickford LB, Erikson ME, Hunt TK, Huang X. Relaxin induces vascular
endothelial growth factor expression and angiogenesis selectively at wound sites.
Wound Repair Regen 8: 361–370, 2000.
542. Unemori EN, Pickford LB, Salles AL, Piercy CE, Grove BH, Erikson ME, Amento EP.
Relaxin induces an extracellular matrix-degrading phenotype in human lung fibroblasts
in vitro and inhibits lung fibrosis in a murine model in vivo. J Clin Invest 98: 2739 –2745, 1996.
543. Van Der Westhuizen ET, Sexton PM, Bathgate RA, Summers RJ. Responses of
GPCR135 to human gene 3 (H3) relaxin in CHO-K1 cells determined by microphysi-
ometry. Ann NY Acad Sci 1041: 332–337, 2005.
544. Van der Westhuizen ET. Molecular characterization of human and mouse relaxin-3
receptors (RXFP3) in recombinant and endogenously expressing cell lines. In: Phar-
macology . Melbourne, Australia: Monash University, 2008, p. 294.
545. Van der Westhuizen ET, Christopoulos A, Sexton PM, Wade JD, Summers RJ. H2
relaxin is a biased ligand relative to H3 relaxin at the relaxin family peptide receptor 3
(RXFP3). Mol Pharmacol 77: 759 –772, 2010.
546. Van der Westhuizen ET, Werry TD, Sexton PM, Summers RJ. The Relaxin Family
Peptide Receptor 3 (RXFP3) activates ERK1/2 through a PKC dependent mechanism.
Mol Pharmacol 71: 1618 –1629, 2007.
547. Vasilenko P, Mead JP, Weidmann JE. Uterine growth-promoting effects of relaxin: a
morphometric and histological analysis. Biol Reprod 35: 987–995, 1986.
548. Velez-Ruiz GA, Sunahara RK. Reconstitution of G protein-coupled receptors into a
model bilayer system: reconstituted high-density lipoprotein particles. Methods Mol
Biol 756: 167–182, 2011.
549. Vinall RL, Tepper CG, Shi XB, Xue LA, Gandour-Edwards R, de Vere White RW. The
R273H p53 mutation can facilitate the androgen-independent growth of LNCaP by a
mechanism that involves H2 relaxin and its cognate receptor LGR7. Oncogene 25:
2082–2093, 2006.
550. Vodstrcil LA, Shynlova O, Verlander JW, Wlodek ME, Parry LJ. Decreased expression
of the rat myometrial relaxin receptor (RXFP1) in late pregnancy is partially mediated
by the presence of the conceptus. Biol Reprod 83: 818 – 824, 2010.
551. Vodstrcil LA, Shynlova O, Westcott K, Laker R, Simpson E, Wlodek ME, Parry LJ.
Progesterone withdrawal, not increased circulating relaxin, mediates the decrease in
myometrial relaxin receptor (RXFP1) expression in late gestation in rats. Biol Reprod
83: 825– 832, 2010.
552. Vodstrcil LA, Wlodek ME, Parry LJ. Effects of uteroplacental restriction on the relaxin-
family receptors, Lgr7 and Lgr8, in the uterus of late pregnant rats. Reprod Fertil Dev
19: 530 –538, 2007.
553. Wade JD, Layden SS, Lambert PF, Kakouris H, Tregear GW. Primate relaxin: synthesis
of gorilla and rhesus monkey relaxins. J Prot Chem 13: 315–321, 1994.
554. Ward DG, Thomas GR, Cronin MJ. Relaxin increases rat heart rate by a direct action
on the cardiac atrium. Biochem Biophys Res Commun 186: 999 –1005, 1992.
555. Way SA, Leng G. Relaxin increases the firing rate of supraoptic neurones and increases
oxytocin secretion in the rat. J Endocrinol 132: 149 –158, 1992.
556. Weisinger RS, Burns P, Eddie LW, Wintour EM. Relaxin alters the plasma osmolality-
arginine vasopressin relationship in the rat. J Endocrinol 137: 505–510, 1993.
557. Weiss G. Relaxin in the male. Biol Reprod 40: 197–200, 1989.
558. Whitmarsh AJ, Davis RJ. Transcription factor AP-1 regulation by mitogen-activated
protein kinase signal transduction pathways. J Mol Med 74: 589 – 607, 1996.
559. Whorton MR, Bokoch MP, Rasmussen SG, Huang B, Zare RN, Kobilka B, Sunahara
RK. A monomeric G protein-coupled receptor isolated in a high-density lipoprotein
particle efficiently activates its G protein. Proc Natl Acad Sci USA 104: 7682–7687, 2007.
560. Wilkinson TN, Bathgate RA. The evolution of the relaxin peptide family and their
receptors. Adv Exp Med Biol 612: 1–13, 2007.
561. Wilkinson TN, Speed TP, Tregear GW, Bathgate RA. Coevolution of the relaxin-like
peptides and their receptors. Ann NY Acad Sci 1041: 534 –539, 2005.
562. Wilkinson TN, Speed TP, Tregear GW, Bathgate RA. Evolution of the relaxin-like
peptide family. BMC Evol Biol 5: 14, 2005.
563. Wilkinson TN, Speed TP, Tregear GW, Bathgate RA. Evolution of the relaxin-like
peptide family: from neuropeptide to reproduction. Ann NY Acad Sci 1041: 530 –533, 2005.
564. Williams EJ, Benyon RC, Trim N, Hadwin R, Grove BH, Arthur MJ, Unemori EN,
Iredale JP. Relaxin inhibits effective collagen deposition by cultured hepatic stellate
cells and decreases rat liver fibrosis in vivo. Gut 49: 577–583, 2001.
565. Willoughby D, Baillie GS, Lynch MJ, Ciruela A, Houslay MD, Cooper DM. Dynamic
regulation, desensitization, and cross-talk in discrete subcellular microdomains during
beta2-adrenoceptor and prostanoid receptor cAMP signaling. J Biol Chem 282:
34235–34249, 2007.
566. Winslow J, Shih A, Laramee G, Bourell J, Stults J, Johnston P. Purification and structure
of human pregnancy relaxin from corpora lutea, serum and plasma. In: 71st Annual
Meeting of the Endocrine Society. 1989, p. 245.
567. Winslow JW, Shih A, Bourell JH, Weiss G, Reed B, Stults JT, Goldsmith LT. Human
seminal relaxin is a product of the same gene as human luteal relaxin. Endocrinology
130: 2660 –2668, 1992.
568. Xu Q, Chakravorty A, Bathgate RA, Dart AM, Du XJ. Relaxin therapy reverses large
artery remodeling and improves arterial compliance in senescent spontaneously hy-
pertensive rats. Hypertension 55: 1260 –1266, 2010.
569. Xu Q, Lekgabe ED, Gao XM, Ming Z, Tregear GW, Dart AM, Bathgate RA, Samuel
CS, Du XJ. Endogenous relaxin does not affect chronic pressure overload-induced
cardiac hypertrophy and fibrosis. Endocrinology 149: 476 – 482, 2008.
570. Yamamoto S, Kwok SC, Greenwood FC, Bryant GG. Relaxin purification from human
placental basal plates. J Clin Endocrinol Metab 52: 601– 604, 1981.
571. Yamamoto T, Davis CG, Brown MS, Schneider WJ, Casey ML, Goldstein JL, Russell
DW. The human LDL receptor: a cysteine-rich protein with multiple Alu sequences in
its mRNA. Cell 39: 27–38, 1984.
572. Yan Y, Scott DJ, Wilkinson TN, Ji J, Tregear GW, Bathgate RA. Identification of the
N-linked glycosylation sites of the human relaxin receptor and effect of glycosylation
on receptor function. Biochemistry 47: 6953– 6968, 2008.
573. Yang S, Rembiesa B, Bullesbach EE, Schwabe C. Relaxin receptors in mice: demon-
stration of ligand binding in symphyseal tissues and uterine membrane fragments.
Endocrinology 130: 179 –185, 1992.
574. Yegorov S, Good-Avila SV, Parry L, Wilson BC. Relaxin family genes in humans and
teleosts. Ann NY Acad Sci 1160: 42– 44, 2009.
575. Ying SY, Becker A, Baird A, Ling N, Ueno N, Esch F, Guillemin R. Type beta trans-
forming growth factor (TGF-beta) is a potent stimulator of the basal secretion of
follicle stimulating hormone (FSH) in a pituitary monolayer system. Biochem Biophys
Res Commun 135: 950 –956, 1986.
576. Yki JH, Wahlstrom T, Seppala M. Immunohistochemical demonstration of relaxin in
the genital tract of men. J Reprod Fertil 69: 693– 695, 1983.
577. Yki JH, Wahlstrom T, Tenhunen A, Koskimies AI, Seppala M. The occurrence of
relaxin in hyperstimulated human preovulatory follicles collected in an in vitro fertil-
ization program. JIn Vitro Fert Embryo Transf 1: 180 –182, 1984.
RELAXIN FAMILY PEPTIDES AND THEIR RECEPTORS
479Physiol Rev VOL 93 JANUARY 2013 www.prv.org
578. Yki-Jarvinen H, Wahlstrom T, Seppala M. Human endometrium contains relaxin that
is progesterone-dependent. Acta Obstet Gynecol Scand 64: 663– 665, 1985.
579. Yki-Jarvinen H, Wahlstrom T, Seppala M. Immunohistochemical demonstration of
relaxin in the genital tract of pregnant and nonpregnant women. J Clin Endocrinol
Metab 57: 451– 454, 1983.
580. Yuan FP, Li X, Lin J, Schwabe C, Bullesbach EE, Rao CV, Lei ZM. The role of RXFP2
in mediating androgen-induced inguinoscrotal testis descent in LH receptor knockout
mice. Reproduction 139: 759 –769, 2010.
581. Zaleski HM, Winn RJ, Jennings RL, Dziuk PJ, Sherwood OD. Effects of relaxin admin-
istration in early gestation or prior to mating on uterine length and fetal survival in gilts.
Biol Reprod 52: 1389 –1394, 1995.
582. Zaleski HM, Winn RJ, Jennings RL, Sherwood OD. Effects of relaxin on lactational
performance in ovariectomized gilts. Biol Reprod 55: 671– 675, 1996.
583. Zarreh-Hoshyari-Khah MR, Einspanier A, Ivell R. Differential splicing and expression
of the relaxin-like factor gene in reproductive tissues of the marmoset monkey (Cal-
lithrix jacchus). Biol Reprod 60: 445– 453, 1999.
584. Zeng L, Akasaki Y, Sato K, Ouchi N, Izumiya Y, Walsh K. Insulin-like 6 is induced by
muscle injury and functions as a regenerative factor. J Biol Chem 285: 36060 –36069,
2010.
585. Zhang J, Qi YF, Geng B, Pan CS, Zhao J, Chen L, Yang J, Chang JK, Tang CS. Effect of
relaxin on myocardial ischemia injury induced by isoproterenol. Peptides 26: 1632–
1639, 2005.
586. Zhang Q, Liu SH, Erikson M, Lewis M, Unemori E. Relaxin activates the MAP kinase
pathway in human endometrial stromal cells. J Cell Biochem 85: 536 –544, 2002.
587. Zhang S, Hughes RA, Bathgate RA, Shabanpoor F, Hossain MA, Lin F, van Lierop B,
Robinson AJ, Wade JD. Role of the intra-A-chain disulfide bond of insulin-like peptide
3 in binding and activation of its receptor, RXFP2. Peptides 31: 1730 –1736, 2010.
588. Zhao L, Roche PJ, Gunnersen JM, Hammond VE, Tregear GW, Wintour EM, Beck F.
Mice without a functional relaxin gene are unable to deliver milk to their pups.
Endocrinology 140: 445– 453, 1999.
589. Zhao L, Samuel CS, Tregear GW, Beck F, Wintour EM. Collagen studies in late
pregnant relaxin null mice. Biol Reprod 63: 697–703, 2000.
590. Zhao S, Fields PA, Sherwood OD. Evidence that relaxin inhibits apoptosis in the cervix
and the vagina during the second half of pregnancy in the rat. Endocrinology 142:
2221–2229, 2001.
591. Zhao S, Kuenzi MJ, Sherwood OD. Monoclonal antibodies specific for rat relaxin. IX.
Evidence that endogenous relaxin promotes growth of the vagina during the second
half of pregnancy in rats. Endocrinology 137: 425– 430, 1996.
592. Zhong M, Murtazina DA, Phillips J, Ku CY, Sanborn BM. Multiple signals regulate
phospholipase CBeta3 in human myometrial cells. Biol Reprod 78: 1007–1017, 2008.
593. Zhou Y, McMaster M, Woo K, Janatpour M, Perry J, Karpanen T, Alitalo K, Damsky C,
Fisher SJ. Vascular endothelial growth factor ligands and receptors that regulate hu-
man cytotrophoblast survival are dysregulated in severe preeclampsia and hemolysis,
elevated liver enzymes, and low platelets syndrome. Am J Pathol 160: 1405–1423,
2002.
594. Zhu J, Kuei C, Sutton S, Kamme F, Yu J, Bonaventure P, Atack J, Lovenberg TW, Liu
C. Identification of the domains in RXFP4 (GPCR142) responsible for the high affinity
binding and agonistic activity of INSL5 at RXFP4 compared with RXFP3 (GPCR135).
Eur J Pharmacol 590: 43–52, 2008.
595. Zimmermann S, Schottler P, Engel W, Adham IM. Mouse Leydig insulin-like (Ley I-L)
gene: structure and expression during testis and ovary development. Mol Reprod Dev
47: 30 –38, 1997.
596. Zimmermann S, Steding G, Emmen JM, Brinkmann AO, Nayernia K, Holstein AF,
Engel W, Adham IM. Targeted disruption of the INSL3 gene causes bilateral cryp-
torchidism. Mol Endocrinol 13: 681– 691, 1999.
BATHGATE ET AL.
480 Physiol Rev VOL 93 JANUARY 2013 www.prv.org
... The effects of E2 in the myometrium include increased responsiveness to contraction-promoting molecules (e.g., oxytocin and prostaglandins), increased gap junction formation, decrease nitric oxide activity, and increased Ca +2 influx (12)(13)(14). The peptide hormone Rln has numerous, roles in the physiology of pregnancy in humans and other species (15)(16)(17). Rln, which is secreted from the corpus luteum and the placenta, mediates its actions through a G-protein coupled receptor, relaxin family peptide receptor 1 (RXFP1) (16). Although Rln is best known for its actions in the cervix (18,19), numerous studies have uncovered diverse roles for Rln in the myometrium (16,17,20,21). ...
... The peptide hormone Rln has numerous, roles in the physiology of pregnancy in humans and other species (15)(16)(17). Rln, which is secreted from the corpus luteum and the placenta, mediates its actions through a G-protein coupled receptor, relaxin family peptide receptor 1 (RXFP1) (16). Although Rln is best known for its actions in the cervix (18,19), numerous studies have uncovered diverse roles for Rln in the myometrium (16,17,20,21). ...
... Rln, which is secreted from the corpus luteum and the placenta, mediates its actions through a G-protein coupled receptor, relaxin family peptide receptor 1 (RXFP1) (16). Although Rln is best known for its actions in the cervix (18,19), numerous studies have uncovered diverse roles for Rln in the myometrium (16,17,20,21). ...
Preprint
Full-text available
A bstract Estradiol (E2) and relaxin (Rln) are steroid and polypeptide hormones, respectively, with important roles in the female reproductive tract, including myometrium. Some actions of Rln, which are mediated by its membrane receptor RXFP1, require or are augmented by E2 signaling through its cognate nuclear steroid receptor, estrogen receptor alpha (ERα). In contrast, other actions of Rln act in opposition to the effects of E2. Here we explore the molecular and genomic mechanisms that underlie the functional interplay between E2 and Rln in the myometrium. We used both ovariectomized female mice and immortalized human myometrial cells expressing wild type or mutant ERα (hTERT-HM-ERα cells). Our results indicate that Rln attenuates the genomic actions and biological effects of estrogen in the myometrium and myometrial cells by reducing phosphorylation ERα on serine 118 (S118). Interestingly, we observed a potent inhibitory effect of Rln on the E2-dependent binding of ERα across the genome. The reduction in ERα binding was associated with changes in the hormone-regulated transcriptome, including a decrease in the E2-dependent expression of neighboring genes. The inhibitory effects of Rln cotreatment on the E2-dependent phosphorylation of ERα required the nuclear dual-specificity phosphatases DUSP1 and DUSP5. Moreover, the inhibitory effects of Rln were reflected in a concomitant inhibition of the E2-dependent contraction of myometrial cells. Collectively, our results identify a pathway that integrates Rln/RXFP1 and E2/ERα signaling, resulting in a convergence of membrane and nuclear signaling pathways to control genomic and biological outcomes.
Article
Full-text available
Human relaxin-2 (H2 relaxin) is a peptide hormone of about 6 kDa, first identified as a reproductive hormone involved in vasoregulation during pregnancy. It has recently attracted strong interest because...
Article
The purpose of the study was to determine whether iv infusion of relaxin would acutely stimulate water drinking in rats and, if it did, whether such drinking is affected by other dipsogenic stimuli or is blocked by centrally administered losartan. iv infusions of human gene 2 relaxin at doses of 25, 40, 55, or 80 μg/kg·h for 1 h induced dose-dependent water drinking in both male and female rats within 15–30 min of commencement of infusions. iv infusion of a nondipsogenic dose of angiotensin II (0.5 μg/h), combined with relaxin (40 μg/kg·h), almost tripled the relaxin-induced water intake. iv infusion of hypertonic (1 m) NaCl did not potentiate relaxin-induced drinking. Intracerebroventricular injection of the angiotensin AT1 antagonist losartan (10 μg) reduced water drinking induced by iv infusion of relaxin. The water drinking induced by iv infusion of relaxin in the rat suggests that blood-borne relaxin may be a dipsogenic hormone. Potentiation of this relaxin-induced drinking by moderate levels of circulating angiotensin II is additional evidence in support of this view. The results also indicate that a central angiotensinergic neural pathway, utilizing AT1 receptors, subserves relaxin-induced drinking.
Article
We have used gene targeting to generate relaxin (rlx)-deficient mice. The majority (15 of 17) of homozygous (rlx−/−) mice are fertile and produce normal litters. However their mammary development is deficient; pups are unable to suckle and die within 24 h of birth unless cross-fostered to a wild-type (rlx+/+) foster mother. The nipples of rlx−/− animals do not enlarge significantly during pregnancy, and their histology retains the appearance of the virgin state. Breast parenchyma is somewhat underdeveloped at term even though milk is produced. Mammary ducts become grossly dilated in these animals. Heterozygous (rlx+/−) mice lactate normally. The interpubic ligament does not relax during pregnancy in rlx−/− mice. Plasma osmolality during late gestation was significantly higher (P < 0.001) in rlx−/− mice than in wild-type controls.
Article
Objective:To determine whether serum hormone profiles are different in nonconceptive and conceptive menstrual cycles after ovulation and before implantation.
Article
It is established that endogenous relaxin promotes the growth and development of the cervix, mammary glands, and nipples in pregnant rats. Additionally, the observation that porcine relaxin promotes growth of the vagina in both nonpregnant and pregnant rats provides indirect evidence that endogenous relaxin may effect growth of the vagina during rat pregnancy. The purpose of this study was to determine whether endogenous relaxin promotes growth of the vagina in pregnant rats. To that end, a monoclonal antibody, specific for rat relaxin, designated MCA1, was used to passively neutralize endogenous relaxin throughout the second half of pregnancy in intact rats. Five milligrams of highly purified MCA1 were injected iv to rats daily from days 12-22 of pregnancy. Controls received either a monoclonal antibody for fluorescein or PBS. The vaginal wet weight, dry weight, length, diameter, inner surface area, DNA content, and percent water content were determined. No differences were found between monoclonal antib...
Article
To elucidate the mechanism of relaxin action, we studied the binding characteristics of human relaxin and its effects on intracellular concentrations of cAMP and tyrosine phosphorylation of cellular proteins in a model system of human cervix, human lower uterine segment fibroblasts. Human relaxin labeled with 125I bound specifically to a single class of high-affinity relaxin binding sites, distinct from insulin receptors, with a mean (±sem) dissociation constant (Kd) of 4.36 ± 1.7 × 10−9 m and a mean of 3220 ± 557 binding sites per cell in human lower uterine segment fibroblasts. Relaxin, in quantities that were shown previously to stimulate intracellular levels of cAMP in other cell types, had no effect on intracellular levels of cAMP in human lower uterine segment fibroblasts even in the presence of the phosphodiesterase inhibitor 3-isobutyl-1-methyl-xanthine (IBMX). Incubation of the cells with relaxin caused a significant increase in tyrosine phosphorylation of a protein with an apparent Mr of approxi...
Article
Experiments were done to examine whether rat relaxin is dipsogenic and whether such dipsogenic effects of rat relaxin are related to time of injection during the light-dark cycle. Female rats were fitted with a chronic intra-cerebro-ventricular (icv) cannula. Rat relaxin (2.5, 5, 10, 25, 50, or 100 ng/2 μl in 0.9% saline) was injected into the right lateral ventricle at either morning (0800–1000 h), afternoon (1400–1600 h), or night (2200–2400 h), and water consumption was measured. Relaxin caused a dose-dependent dipsogenesis at doses ≥ 5 ng, but the sensitivity and magnitude of the response varied with the photoperiod. Water consumption was smallest (3.5 ± 0.7 ml at 50 ng) and least sensitive (minimal effective dose at 25 ng) in the afternoon and maximal (17.7 ± 2.3 ml at 50 ng) and most sensitive (minimal effective dose 5 ng) at night. The latency from injection to drinking was 55.8 ± 10.4 sec (mean ± sem) and did not vary significantly with either the dose or time of day. A second set of experiments w...
Article
We have prepared a polyclonal antibody (AntiF) against a synthetic peptide comprising residues 19-29 of the rat FSH receptor. The specificity of this antibody was documented using human embryonic kidney (293) cells and stable transfectants of 293 cells expressing the recombinant LH/CG and FSH receptors. The data presented show that AntiF inhibits the binding of FSH, but not that of LH/CG, to their cognate receptors. AntiF also recognizes a specific protein(s) representing the FSH receptor in Western blots or by immunoprecipitation of cells transfected with the FSH receptor. This protein(s) is not present in untransfected 293 cells or in 293 cells permanently transfected with the LH/CG receptor. Finally, addition of the 19-29 peptide prevents immunoprecipitation of the FSH receptor by AntiF, whereas an unrelated peptide corresponding to residues 637-647 has no effect. In Western blots of 293 cells transfected with the FSH receptor, AntiF reveals the recombinant receptor as a heterogenous glycoprotein with ...
Article
Experiments were performed to determine whether neutralization of relaxin in the brain, by injecting monoclonal antibodies to rat relaxin into the ventricular system of the brain, affected either the timing or the processes of birth in rats. Pregnant rats were injected daily through a chronically implanted intracerebroventricular cannula either with a specific monoclonal antibody raised against rat relaxin from days 12–22 of gestation or with an antibody raised against fluorescein as a control. The rats were watched closely from the afternoon of day 20 of pregnancy, and the process of birth was observed. No sign of dystocia was observed in any of the rats in the experiment. Neutralization of endogenous relaxin caused a significant decrease in the length of gestation (505.4 ± 3.1 h) compared with that in rats treated with PBS (524.6 ± 0.5 h) or that in rats treated with a nonspecific antibody (525.9 ± 0.7 h). The time to the onset of delivery was also shorter in the relaxin-neutralized group (507.8 ± 1.1 h) compared with that in either PBS-treated (526.5 ± 0.6 h) or fluorescein antibody-treated (525.3 ± 0.7 h) animals. In contrast, there was no significant effect of the relaxin antibody on length of straining, duration of parturition, delivery interval, live birth rate, or body weight of the neonates. Premature delivery in the relaxin-neutralized group was accompanied by a 24-h advance in the fall in plasma progesterone. These data support the hypothesis that there may be a central relaxin system that is independent of the peripheral relaxin system. Central relaxin may have a significant physiological role on the timing of pregnancy in the rat, but does not affect the course of labor once it has started.